Next Article in Journal
Multiomics Analysis of Endocytosis upon HBV Infection and Identification of SCAMP1 as a Novel Host Restriction Factor against HBV Replication
Next Article in Special Issue
BnERF114.A1, a Rapeseed Gene Encoding APETALA2/ETHYLENE RESPONSE FACTOR, Regulates Plant Architecture through Auxin Accumulation in the Apex in Arabidopsis
Previous Article in Journal
Capsaicin and Zinc Promote Glucose Uptake in C2C12 Skeletal Muscle Cells through a Common Calcium Signalling Pathway
Previous Article in Special Issue
Simultaneous Improvement of Grain Yield and Quality through Manipulating Two Type C G Protein Gamma Subunits in Rice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Improvement and Re-Evolution of Tetraploid Wheat for Global Environmental Challenge and Diversity Consumption Demand

1
Australia-China Joint Centre for Wheat Improvement, College of Science, Health, Engineering and Education, Murdoch University, Perth, WA 6150, Australia
2
Triticeae Research Institute, Sichuan Agricultural University, Chengdu 611130, China
3
Crop Research Institute, Sichuan Academy of Agricultural Sciences (SAAS), Chengdu 610066, China
4
College of Agronomy, Qingdao Agricultural University, Qingdao 266109, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(4), 2206; https://doi.org/10.3390/ijms23042206
Submission received: 24 December 2021 / Revised: 21 January 2022 / Accepted: 24 January 2022 / Published: 17 February 2022
(This article belongs to the Special Issue Yield Traits and Their Genetic Pathway in Crop)

Abstract

:
Allotetraploid durum wheat is the second most widely cultivated wheat, following hexaploid bread wheat, and is one of the major protein and calorie sources of the human diet. However, durum wheat is encountered with a severe grain yield bottleneck due to the erosion of genetic diversity stemming from long-term domestication and especially modern breeding programs. The improvement of yield and grain quality of durum wheat is crucial when confronted with the increasing global population, changing climate environments, and the non-ignorable increasing incidence of wheat-related disorders. This review summarized the domestication and evolution process and discussed the durum wheat re-evolution attempts performed by global researchers using diploid einkorn, tetraploid emmer wheat, hexaploid wheat (particularly the D-subgenome), etc. In addition, the re-evolution of durum wheat would be promoted by the genetic enrichment process, which could diversify allelic combinations through enhancing chromosome recombination (pentaploid hybridization or pairing of homologous chromosomes gene Ph mutant line induced homoeologous recombination) and environmental adaptability via alien introgressive genes (wide cross or distant hybridization followed by embryo rescue), and modifying target genes or traits by molecular approaches, such as CRISPR/Cas9 or RNA interference (RNAi). A brief discussion of the future perspectives for exploring germplasm for the modern improvement and re-evolution of durum wheat is included.

1. Introduction

Wheat is one of the main cereal crops in the world and a major source of carbohydrates and proteins in the human diet [1,2,3]. An adequate supply of wheat and its processed products (i.e., flour, bread, pasta, biscuit, bear) [4,5,6] can not only ensure food security and social stability but also enrich dietary diversity and avoid malnutrition [2,7,8]. The forecasted world population of 9.6 billion people in 2050 requires increasing wheat production by 60% in the next 30 years [9,10]. Moreover, the ever-changing climate brings great challenges to the sustainability of global wheat production [11,12,13,14,15,16]. Hence, the improvement of wheat on yield, quality, biotic resistance, abiotic tolerance, and diversity of the processed products has become a paramount task to cope with global climate change and to meet diverse consumption demands.
Wheat cultivars are usually divided into two groups, including tetraploid durum wheat (Triticum turgidum L. concv. durum, 2n = 4x = 28, AABB) and hexaploid bread wheat (Triticum aestivum L., 2n = 6x = 42, AABBDD) [17]. Allohexaploid wheat was derived from the spontaneous hybridization between tetraploid wheat and diploid Aegilops tauschii Cosson (2n = 2x = 14, DD) [18,19,20]. The greatly improved plasticity and adaptability of bread wheat conferred by the addition of the D genomes promotes it as one of the major cultivated crops in the world, accounting for approximately 95% of global wheat production [21,22]. Compared to bread wheat, durum wheat only accounts for the other 5%. Nevertheless, it is widely cultivated in many countries, including Turkey, Canada, Algeria, Italy, India, Australia, etc., with approximately 42.7 million tons of global production in 2020–2021 [23,24]. Compared to the other subspecies of tetraploid wheat, durum wheat is predominant in cultivation and food production due to its highly desirable phenotypes and diversified products for human consumption [24,25,26].
So far, many studies have reported the improvement of bread wheat performance through the introgression of the genetic variations of tetraploid wheat. Meanwhile, a few research works have reported the re-evolution of durum wheat through utilizing genes from other tetraploid subspecies or hexaploid wheat (particularly mediating the introgression of D genomes) [22,27,28]. This review summarizes the origin, domestication, evolution, and breeding strategies and achievements of tetraploid wheat, especially durum wheat. The updated information can shed some light on the future improvement of durum wheat to cope with global environmental challenges, meet the diversity consumption demands, and reduce wheat-related disorders in humans.

2. Origin, Domestication, and Evolution of Tetraploid Wheat

2.1. Origin Process of Tetraploid Wheat

Tetraploid wheat is constituted by two groups: Timopheevi group wheat (Triticum timopheevi Zhuk., 2n = 4x = 28, AuAuBspBsp/AuAuGG) and Turgidum group wheat (Triticum turgidum L., 2n = 4x = 28, AuAuBB) (Table 1), according to the Biosystematics of Triticeae updated by Yen and Yang [17]. The origin of allotetraploid wheat can be traced back to presumably 0.3–0.5 million years before present (BP) in or near the oak-pistachio woodland belt, also called Near Eastern Fertile Crescent (Figure 1) [29,30,31].
Wild emmer wheat (T. turgidum var. dicoccoides, 2n = 4x = 28, AuAuBB) with brittle rachis (Br) and hulled seeds originated from the natural hybridization between two diploid ancestors, including wild einkorn wheat Triticum urartu (Triticum monococcum ssp. urartu, 2n = 2x = 14, AuAu) that provided the A genome [32,33], and a wild species of the Aegilops genus provided the B genome [34]. The donor of the A-genome for all tetraploid and hexaploid wheat was proposed to be another wild einkorn wheat Triticum boeoticum (Triticum monococcum ssp. aegilopoides, 2n = 2x = 14, AbAb), the ancestor of cultivated einkorn Triticum monococcum (2n = 2x = 14, AmAm), according to the early cytogenetic investigation, whereas it was proved to be the T. urartu later [31,32,33,35,36,37,38,39]. Although the origin of the B genome is still uncertain, immense geographical, morphological, cytological, genetic, biochemical, and molecular evidence has been accumulated, suggesting Aegilops speltoides (SS) as the contributor of the B genome [17,34,40,41,42,43,44]. However, recent research indicated that Ae. speltoides was not the direct progenitor of the B genome [45]. The comparison of genome sequences of five Aegilops Sitopsis species showed that Ae. speltoides and the B-subgenome diverged about 4.49 million years BP, which was much earlier than the speciation of tetraploid emmer wheat [45]. The long evolution history after tetraploid wheat generation may result in the high divergence of B genomes between hexaploid and diploid donors, making it difficult to deduce the precise donor/s [46,47].
T. araraticum (T. timopheevi var. araraticum, 2n = 4x = 28, AuAuBspBsp) was generated in another independent polyploidization event between T. urartu and other subspecies of Ae. speltoides (BspBsp/GG), representing the wild forms of timopheevi wheat [41,48,49,50]. It is indistinguishable from T. dicoccoides in morphology but different in genomic constitution, therefore showing irregular meiosis and high infertility when crossed with T. turgidum group wheat or hexaploid wheat [17,29].
Wild emmer is native to the Fertile Crescent and is about 0.36 million years old [51]. It was at 19 thousand years BP that hunter–gatherers started to collect and use wild emmer wheat based on the archaeological evidence found at Ohalo II located on the southwestern shore of the Sea of Galilee, Israel [52,53]. The first spikelet of wild emmer wheat was isolated by T. Kotschy from specimens of wild barley (Hordeum spontaneum K.) in 1855 and recognized as wild wheat by Kornicke until 1873 [20,54]. In 1906, the wild emmer wheat was rediscovered in nature by Aaron Aaronsohn near Rosh Pinna, eastern Galilee [55]. According to the literature, this species is still distributed in present-day in the Jordan valley, Southeastern Turkey, Eastern Iraq, and Western Iran [29,54]. It is usually growing in various geological conditions, such as basalt areas, hard limestone bedrocks, Terra Rossa soils, and grass and woodland of the hill country, with altitudes ranging from about 150 m below sea level to 1800 m above sea level [29,54]. The complicated grow habitats might promote the genetic differentiation and evolution of wild emmer wheat.

2.2. Domestication and Evolution Process

The domestication of wild emmer wheat started about 10 thousand years BP (Figure 1), which symbolized the first and significant evolution intervention by humans and raised the Neolithic revolution [20,56]. Consequently, two cultivated tetraploid wheat, T. turgidum concv. dicoccon (2n = 4x = 28, AuAuBB) and T. timopheevi concv. timopheevi (2n = 4x = 28, AuAuBspBsp), were domesticated from T. dicoccoides and T. araraticum, respectively [29,57,58,59,60,61,62]. Since the birth of domesticated emmer wheat in the Fertile Crescent, T. dicoccon has been widely spread from farming villages in Pre-Pottery Neolithic B (PPNB) to the Mediterranean region and has become one of the most prominent crops for almost 6 thousand years [35,63]. Around 9 thousand years BP, allohexaploid Spelt wheat (Triticum spelt, 2n = 6x = 42, AuAuBBDD) was produced through the natural hybridization between T. dicoccon and Ae. tauschii [18,19,26]. In addition, the domesticated T. timopheevii naturally hybridized with T. boeoticum gave rise to the hexaploid Zhukoyskyi wheat (Triticum zhukovskyi, 2n = 6x = 42, AuAuBspBspAbAb) in Transcaucasia [26,64]. About 8.5 thousand years BP, the currently most cultivated durum wheat (T. turgidum concv. durum) with a tough rachis and free-threshing seeds was produced from natural mutation of T. dicoccon in the Eastern Mediterranean region (Figure 1) [52,63]. The subsequential polyploidization between domesticated emmer T. dicoccon (or cultivated T. durum) and Ae. tauschii Coss. gave rise to bread wheat [20,26]. Additionally, some research indicated that bread wheat was produced by the mutation of Triticum spelt [20,26].
According to the results of morphological and molecular evidence, two distinct populations of T. dicoccoides have been identified in Southern and Northern Levantine Corridor sites, resulting in dissident opinions on the geography of domestication [52,63,65,66]. The amplified fragment length polymorphism (AFLP) marker analysis was initially used to compare the wild and domesticated emmer wheat and found that most of the wild emmer lines collected from the Karacadag mountain of southeastern Turkey were more related to domesticated emmer and durum wheat, laying the foundation of monophyletic origin opinion [67]. The re-analysis of the AFLP data through principal coordinate analysis (PCA) supported the previous result of Özkan [68]. Further research proposed an independent domestication event of wild emmer wheat that occurred in Karacadag mountain and somewhere else in southeastern Turkey, inferring from the result of chloroplast DNA fingerprinting [65]. Subsequently, larger populations were used to reconsider the domestication geography of tetraploid wheat [69]. The authors suggested that the central-eastern race was the progenitor of domesticated emmer wheat since it played a key role during domestication. They also declared disagreement evidence on the domestication origin site based on the data of chloroplast DNA (Kartal-Karadag mountain) and AFLP analysis (Karacadag range) [69]. In 2007, more tetraploid accessions were investigated by Luo et al. [63], who agreed with the monophyletic domestication in Northern Levant followed by subsequent hybridization and introgression from wild to domesticated emmer in the southern Levant. Another possibility is that T. dicoccoides was independently domesticated in the Northern and Southern Levant [63]. Archaeobotanical findings are consistent with the polycentric domestication model of wild emmer wheat across the Levant [52]. Following this research, Özkan et al. [54] suggested that there was a pre-adaption wild emmer race spreading to several locations in the Fertile Crescent before being domesticated. So far, the theory of multiple site-independent domestications of T. dicoccoides across the Levant might be more realistic [48]. Accordingly, multiple genes and traits were transferred and involved in selection and evolution through numerous hybridizations and mutations. Consequently, polymorphic populations, rather than single genotypes, evolved from the domesticated emmer wheat, such as Rivet wheat (T. turgidum concv. turgidum), Polish wheat (T. turgidum concv. polonicum), and Persian wheat (T. turgidum concv. carthlicum) [17,58].

2.3. Variations of Major Traits during Domestication and Evolution

Crop domestication aims to meet human needs and environmental conditions. Wild wheat, barley, and rye were preferred over the other cereals for ancient people, attributed to the large spikes with large and heavy grains [29,52]. Furthermore, the domestication of wild emmer was much better than barley since the former has more and larger grains [14]. The purposeful selection of domestication accompanied by the stress of specific agroecological environments, natural hybridization, and mutation has led to the ‘domestication syndrome,’ including genomic, morphologic, and phenotypic variations. The evolution from brittle rachis, tough glumes, and hulled seeds to non-shattering, soft glumes, and free-threshing might be the primary domestication targets of wild emmer wheat, as the processes facilitated harvesting by the primitive farmer [31,48].

2.3.1. Changes from Brittle to Non-Brittle Rachis

The brittle rachis (Br) promotes wild tetraploid wheat to disperse seeds freely at maturity through developing abscission of fracture zone at the joint of articulation of the spikelet and rachis, which results in high yield reduction [70]. However, the tough rachis of domesticated forms of emmer wheat suppressed seed dispersal and self-planting and made grain harvesting feasible. Hence, the target transformation of spikes from Br to non-Br was constantly conducted by early farmers for more than one thousand years, symbolizing the first trait of domestication in wheat [56,71,72]. This qualitative trait modification is critical for the origin of agriculture and sedentary societies [20].
The dominant Br genes have been mapped on the short arms of the chromosomes of homologous group 3 in both tetraploid and hexaploid wheat [73,74,75,76] and their key roles have been identified [77,78,79]. At first, a single dominant gene, Br1 (Br-A1), was identified that controlled rachis fragility in feral or semi-wild hexaploid wheat from Tibet [80]. This gene was mapped on the short arm of the 3D chromosome later [74]. Two other dominant genes, Br2 (Br-A2) and Br3 (Br-A3), controlled the Br phenotypes in wild emmer wheat and were localized on the short arms of 3A and 3B chromosomes, respectively [59,81]. Three types of rachis fragility, semi-wild wheat type, spelta-type, and tough rachis type, were controlled by the interaction of these three genes [20]. The haplotype analysis of Br1-A and Br1-B genes confirmed the common origin of the cultivated Turgidum group wheat and indicated a separate domestication event of T. timopheevii [61]. The recessive alleles from mutant Br genes conferred non-shattering spikes to domesticated emmer, facilitating harvest [20,26].

2.3.2. Variations from Non-Free Threshing to Free Threshing

The grains of both wild and domesticated tetraploid wheat are wrapped with tough glumes (Tg), consequently making them hard to thresh [82]. On the contrary, all cultivated tetraploid wheat is characterized by soft glumes and free-threshing (non-hulled seeds) [31,83]. The evolution of the free-threshing trait in durum wheat was mainly attributed to the mutation of the semi-dominant Tg gene on the short arms of 2A and 2B and another critical domestication dominant gene, Q (primitive allele as q), located on the long arm of the 5A chromosome [83,84,85,86,87,88,89,90,91,92,93]. Compared with the Q gene, Tg showed a more pronounced effect on the threshability of seeds [85]. All cultivated tetraploid wheat that display the free-threshing phenotype should carry QQtgtg combinations. Therefore, the QQTgTg genotype might be concealed in the domestication process of emmer wheat [26].
The Q gene originated from a single amino acid mutation from the wild type of the q allele and encoded a transcription factor belonging to the APETALA2 (AP2) family [89,94]. The variations between Q and q alleles took place in the conserved coding region, from A to G at position 985 and from C to T at position 1254. The Q gene enhanced protein dimerization activity relative to q and reduced the binding and degradation of microRNA172, resulting in higher transcription levels, which lead to a compact spike with a free-threshing phenotype [95,96]. The fragile rachides of wild wheat and the tough rachis of the semi-dominant mutant indicate the other role of the Q gene in controlling the trait of brittle rachis [97]. In addition, dosage effects of Q and q alleles were observed on the spike phenotype [98]. The loss-of-function mutant of Q and q genes led to a speltoid-like spike and non-free-threshing phenotype [99]. Considering the pleiotropic effects of the Q gene on rachis fragility, glume tenacity, threshability, spike architecture, plant height, and flowering time, it could be considered a gain-of-function mutation and recognized as the most important domestication gene [83,94,96,99,100,101].

2.3.3. Other Qualitative Domestication Traits Involved

During the domestication process, many other traits were co-selected, including plant height, tiller number, flowering/heading time, phenotypes of the spike, spikelet, seed, etc. [31,48,71,83,96,102,103,104,105,106,107]. For example, spring wheat lacked vernalization and specific photoperiod requirements were domesticated from winter wheat for good adaptation to the prevailing environmental conditions in the Fertile Crescent region [20]. The evolution of this adaptation model plays a key role in the post-domestication for the adaptation of temperate cereals [108]. In tetraploid wheat, the QTLs related to flowering/heading time were mapped on 2A, 4B, 5A, and 6B [71]. Among them, the wild allele for the QTL on 5A is associated with late-anthesis of wild emmer, whereas the other three QTL are responsible for early flowering [71]. The QTL mapped on 2A is present in a collinear position with the photoperiod response (Ppd) genes (2AS, 2BS, 2DS) [48]. Plant height affects lodging and consequently grain yield and quality traits [109]. Modern dwarf wheat was attributed to the reduced height gene Rht1 (Rht-B1b) (located on 4B chromosomes), which can substantially reduce the height of wheat, prevent plant lodging, and increase the harvest index [109,110,111]. Many dwarfing genes have been mapped on the AB genomes of wheat and classified into the phytohormone gibberellin (GA)-insensitive (i.e., Rht1, Rht3 (Rht-B1c), and Rht11 (RhtB1e)) and GA-sensitive (i.e., Rht9, Rht18, and Rht24) groups based upon the plant response to bioactive GA [112]. Recently, a recessive semi-dwarfing gene Rht-dp was fine mapped in a Polish wheat line, which was subsequently identified as the same gene of Green Revolution contributor Rht1 [112,113]. These evolved qualitative and quantitative traits under the driving force of domestication had a substantial impact on the early cultivation of cereal crops and the rise of modern agriculture [99].

2.4. Importance and Breeding Challenge of Durum Wheat

After thousands of years of empirical selection and breeding, modern cultivated tetraploid wheat and hexaploid wheat were formed and widely cultivated, accompanying morphological, phenotypical, and physiological changes [31,48,71]. The domestication and evolution led to modern cultivated tetraploid and hexaploid wheat superseding einkorn, emmer, and spelt wheat, which are treated as relic crops with minor economic importance [1,21]. Einkorn was used at the beginning of agriculture in Europe, whereas its cultivation started to decline in the Bronze Age, probably due to the emergence of free-threshing wheat [60,114]. Although the free-threshing form of einkorn was discovered as well, its cultivation area was limited since the soft glume was associated with the reduction of ear length and grain yield [60]. The cultivation of the tetraploid T. timopheevii wheat was limited as well because of its hard threshing trait and infertility of hybrids when crossed with T. turdidum or T. aestivum [49,61]. Among the different tetraploid wheat, durum wheat is currently the most widely cultivated subspecies, particularly in the Mediterranean region, which accounts for more than half of the worldwide growing region of durum wheat [25]. There are probably a couple of reasons for the durum wheat adaptation. First, along with the einkorn wheat declination, durum wheat had almost completely substituted the wild and domesticated emmer wheat based on its ideal phenotypes (tough rachis, soft glumes, free-threshing ability, large seeds, and short dormancy, etc.) at the beginning of the 20th century [115,116]. Exceptionally, T. dicoccum is still the main food resource for Romans in Italy [115]. Second, durum wheat can be used for diverse food products in terms of its specific characterizations [117,118]. Pasta is the most common end product of durum wheat and supplied 15.8 million tons in the world in 2019 [119]. In addition, the evolution of durum wheat has gone through a long history from wild emmer, domesticated emmer, cultivated durum landraces, to modern durum cultivars, accompanying allopolyploidy from diploid to tetraploid [24,120], which makes durum wheat an excellent research model for the evolution of allopolyploid speciation, adaptation and domestication in plants [121]. Moreover, durum wheat is cross-compatible and inter-fertile with bread wheat since they share the same AB genomes [122].
However, undesirable domestication syndrome of cultivated tetraploid wheat exists. Accompanied with domestication and especially breeding programs of modern agriculture with limited primal parents, the genetic diversity of current cultivars dramatically declined, which in turn led modern cultivars to be susceptible and vulnerable to biotic or abiotic stress [120,123]. Haudry et al. [124] revealed that during domestication, the genetic diversity in the cultivated forms was reduced by 84% in durum wheat and 69% in bread wheat. Genomic comparative analysis between Svevo (modern durum wheat cultivar) and Zavitan (wild emmer wheat) revealed that regions exhibiting strong signatures of genetic divergence (associated with domestication and breeding) were widespread in the genome, with several major diversity losses in the pericentromeric regions [120]. They also identified a high-cadmium accumulation allele widespread among durum cultivars but undetected in wild emmer accessions, which was caused by a non-functional variant of a metal transporter TdHMA3-B1 encoding gene [120]. Therefore, it is a paramount task to reinforce the genetic diversity of durum wheat and improve its yield, grain quality, biotic resistance, and abiotic tolerance to cope with the rapidly growing population and the challenges of global warming [9,10,11,12,13,14,15,16,125].

3. Improvement and Re-Evolution of Durum Wheat

3.1. Strategies and Approaches

Since wild emmer wheat contains numerous favorable genes and shows complete genomic compatibility with durum wheat, it can be directly used to enrich the genetic diversity of durum wheat [126,127]. Aegilops species also contain numerous useful genes [128]. Compared to hexaploid wheat with the addition of the D genomes, durum wheat has limited genetic variations and allelic combinations, resulting in narrow adaptability to different photoperiods, vernalization, and soil conditions. These characterizations inserted a negative impact on yield potential, end-use quality, and capacities to make various food products [21,129,130]. Considering the great effects of the D genome of Ae. tauschii, researchers also proposed to introduce the D genome to re-evolve durum wheat through hybridizing with hexaploid wheat that can readily produce hybrids [21,22,28,122,129]. In diploid wheat, the useful traits from the sources of einkorn wheat other than the D genome can also be integrated into durum wheat smoothly [131,132].
To date, a wide range of approaches have been used for the improvement of durum wheat, as summarized in Table 2. Hybridization directly with diploid einkorn, wild emmer, and hexaploid wheat has been the most traditional breeding method [20,122,133]. Similarly, inactivation of the pairing of homologous chromosomes gene Ph located on the 5B chromosome can induce pairing of homoeologous chromosomes for mediating genetic introgression in wheat-alien hybrid combinations. The most common mutants include durum wheat ph1c and bread wheat ph1b [134,135,136]. Moreover, Joppa developed a set of disomic substitution lines, in which the chromosome pairs of tetraploid wheat Langdon were replaced by the homoeologous D-chromosome pair from Chinese Spring (CS, a hexaploid wheat landrace) [137,138]. These substitution lines have been widely used to integrate the D-genome into durum wheat cultivars. On the other hand, with the development of modern biotechnology, speedy and effective breeding of durum wheat can be achieved through molecular marker-assisted selection (MAS), genome sequencing-assisted breeding, and target gene modification, such as RNA interference (RNAi) and CRISPR-Cas9, leading to many durum wheat lines with improved agronomic traits, protein quality, biotic resistance, abiotic tolerance, and reduced or eliminated triggering factors of wheat plant disorders [139,140,141,142,143,144].

3.2. Major Achievements

3.2.1. Crop Yield Potential

Crop yield is the main goal of durum wheat breeding associated with multiple component traits, such as plant height, tiller number, spike number, spike length, spikelet number per spike, kernel number per spike, and thousand-grain weight. Semidwarf cultivars of durum wheat were generated by introducing the Rht1 (Rht-B1b) gene from Norin 10, leading to a significantly increased harvest index [145]. Compared to durum wheat, bread wheat can utilize more dwarf genes, such as the homoeologous gene Rht2 (Rht-D1b) on the 4D chromosome [109]. Recent research revealed that Rht1 could reduce plant height by generating an N-terminal truncated DELLA protein, the key repressors of the GA signaling pathway and plant growth, through tissue-specific translational reinitiating, which has no effect on the dormancy of seeds [189]. In durum, two dwarfing genes, Rht14 and Rht15, have been introgressed into cultivar Langdon through crossing with dwarf lines [146,147]. However, the introgressed lines showed a decrease in yield component traits. By crossing durum wheat with hexaploid wheat, some tetraploid lines were developed with largely increased spike length, spikelet per spike, grain number per spike, and thousand-grain weight through introgressing the genome of hexaploid wheat using inter-ploidy hybridization [148]. Recently, pentaploid hybridization has also been used to cross durum wheat with its nascent synthetic hexaploid wheat and developed a durum wheat—Ae. tauschii Coss. 4D (4B) disomic substitution line YL-443 [149]. Compared to the tetraploid parent, the developed line showed a larger spike with an increased number of spikelets and florets per spike by 36.3 and 75.9%, respectively [149]. In addition, the kernel size of durum wheat could be significantly increased by knocking out the grain weight gene GW2 through RNAi technology, leading to increased grain starch content by 10–40% [150].

3.2.2. Grain Quality

The grain quality of durum wheat is usually related to kernel hardness, grain protein content, dough tenacity and extensibility, gluten strength, yellow pigment concentration, etc. [117,190]. The kernel texture of hexaploid wheat is controlled by the Hardness locus located on the short arm of the 5D chromosome, which contains two puroindoline genes, Pina-D1 and Pinb-D1, and the grain softness protein gene Gsp-D1 [191]. In diploid wheat, these three genes of the Hardness locus are harbored on 5A or 5B chromosomes, whereas the Pin genes were deleted from these two chromosomes in durum wheat, resulting in the loss of the softness-conferring PIN proteins and consequently a generally hard texture [192]. The soft kernel character of hexaploid wheat contributed to the reintroduction of the Pin genes from Ae. tauschii Coss. [193]. Nevertheless, soft kernel texture of durum wheat with improved baking quality was observed through introducing the Hardness locus from the 5D chromosome of bread wheat, of which the hybridization was mediated by ph1b or ph1c lines [28,151,152]. By crossing the Pina-expressing transgenic line with high molecular weight (HMW) glutenin subunit 1A × 1-expressing transgenic line of durum wheat, a few 1A × 1 and Pina co-expressing lines were generated, showing a significantly improved breadmaking quality and unaffected pasta-making potential [153]. Compared to the transgenic lines, which only express 1A × 1 or Pina, the 1A × 1 compensated for the detrimental effect of the Pina gene (reduce water absorption and damage starch) in co-expressing lines [153]. High protein content and pure durum semolina are strictly required, especially in Italy, France, and Greece, for manufacturing pasta or some bakery products [190]. The HMW glutenin encoding locus Glu-D1 was transferred from 1D of common wheat into 1A of the recipient durum wheat and greatly improved bread-making quality of the latter [28,154,155,156,157,158,159]. It was concluded that the Dx2 + Dy12 was superior to another Glu-D1 allelic variant Dx5 + Dy10 on bread-making quality in durum wheat, as the latter produced excessively strong and inelastic doughs [194]. Moreover, the grain protein content gene Gpc-B1, a NAC (NAM, ATAF1/2, CUC2) transcription factor for high grain protein, zinc, and iron contents, was successfully introgressed into two high-yielding but lower protein Canadian durum wheat lines [126,160].
Yellow pigment concentration represents the carotenoid accumulation in the kernels of durum wheat and has been regarded as a source of important nutrients or antioxidant compounds [195]. A DNA marker PSY-1SSR for yellow pigment concentration QTL, Qyp.macs-7A, was developed and used for marker assistant selection in durum wheat breeding [196]. Multiple mutant combinations of the β-carotene synthetic key inhibitor genes, lycopene ε-cyclase (LCYe) and β-carotene hydroxylase 2 (HYD2), were isolated by Yu et al. [161] through Targeting Induced Local Lesions in Genomes (TILLING), which significantly increased β-carotene in the endosperm for biofortification of provitamin A in durum wheat.
In addition, pre-harvest sprouting could significantly degrade the end-use quality of durum wheat and cause a huge reduction in yield, occurring especially in cool and moist conditions after maturity [162]. A major QTL, Qphs.sicau-3B.1, was introgressed from the long arm of the 3B chromosome in hexaploid wheat T. spelta into durum wheat, conferring its high resistance to pre-harvest sprouting [162]. Sequence characterized amplified region (SCAR) markers were also developed for tracking and breeding durum wheat cultivars with resistance to pre-harvest sprouting, higher yield, and good end-use quality [162].

3.2.3. Biotic Resistance

Wheat grows in diverse geographical regions and environments under different production systems, making it exposed to various biological pathogens that might cause a huge loss of yield or decrease in grain quality [197]. Common diseases include leaf rust (LR), stripe rust (YR), stem rust (SR), powdery mildew (PM), fusarium head blight (FHB), etc. [197]. Through crossing with resistant durum wheat or hexaploid wheat, the Fhb1, Pm13, and Lr19 (tightly linked with yellow pigment in endosperm) genes and some disease-resistance QTLs were successfully transferred into durum wheat [156,163,164,171,172,173,174]. It was reported that some durum wheat lines carried multiple improved traits, such as resistance to leaf rust (Lr14) and tan spot (races 4 and 6), higher grain yield, and higher grain protein content, through crossing with T. araraticum, T. dicoccoides, and Ae. speltoides [165]. By introducing resistance genes Lr34/Yr18/Sr57/Pm38/Ltn1 from hexaploid wheat, durum wheat was conferred with a robust seedling resistance to leaf rust, stripe rust, and powdery mildew diseases [166]. The resistant genes of stem rust (Sr22) and stripe rust (none specify) introgressed from the A genome of diploids enhanced the disease resistance of durum wheat [167,168].
Synthetic wheat is another way to introduce biotic resistance genes to durum wheat. The synthesized resistant AABBAA amphiploid of T. durum and T. boeoticum, T. monococcum, or T. urartu are useful genetic resources for stripe rust, leaf rust, and powdery mildew resistance, which could be transferred into cultivated durum wheat [168,169,170]. Previous studies indicated that the stripe rust resistance gene Yr28 of durum wheat-Ae. tauschii 4D (4B) disomic substitution line could be fully expressed in the tetraploid background, whereas partially expressed in the hexaploid background, which can be used for improving stripe rust resistance in durum wheat breeding [149]. These results demonstrate the successful synthetic methodology in durum wheat breeding programs.
In addition, advanced durum wheat lines with promising FHB resistance were screened out through treatment with 5-azacytidine (DNA methylation inhibitor), which might be contributed by a significant number of differentially expressed genes related to the biosynthesis of secondary metabolites, MAPK signaling, photosynthesis, etc. [175]. Hessian fly is a severe pest of winter wheat, resulting in a reduction of grain and forage production through stunting and killing vegetative tillers, preventing spike developing, and reducing grain filling [198]. Increased resistance to Hessian fly together with superior agro-phenological traits have been obtained in recombinant inbred durum wheat lines [176].

3.2.4. Abiotic Tolerance

Environmental stress, such as drought, high or low temperature, and soil salinity, pose a great threat to the production of durum wheat. Enhancing the environmental adaptability of durum wheat has great potential for increasing its yield and quality. The thermotolerance of durum wheat could be improved by inducing the expression of heat shock proteins (HSPs) [177,178]. By introgression of alien genes from wild emmer, durum wheat displayed a significant increase in the root-to-shoot ratio for enhancing water stress tolerance [179]. Overexpression of TdPIP2;1, a wheat aquaporin gene, could significantly enhance drought and salt tolerance in transgenic lines of durum wheat by reducing excessive water evaporation from leaves in response to water deficit [180]. The ascorbic acid (AsA) treatment could also be useful in enhancing the salt tolerance of durum wheat, whereas it might not be suitable for field use [181]. In bread wheat, the major Al3+ tolerance Kna1 locus and TaALMT1 gene are mapped on the 4D chromosome [199,200]. The ph1-mediated pentaploid hybridization and intraspecific crosses have been used to transfer these genes or loci from bread wheat and Langdon-CS 4D (4B) substitution line, respectively, into durum wheat for developing advanced tolerant lines [182,183,184]. Other than the D genome resistant sources, a wide range of crosses with subspecies of the turgidum group of wheat (emmer, durum, carthlicum, turgidum, turanicum, polonicum) also improved durum wheat’s salt tolerance [185,186].

3.2.5. Wheat-Related Disorders in Humans

During the last decades, an increasing incidence of wheat-related disorders has been reported in human beings worldwide, such as celiac disease (CD), wheat allergy (WA), wheat-dependent exercise-induced anaphylaxis (WDEIA), and non-celiac wheat sensitivity (NCWS) [201,202,203]. Currently, there are 28 allergens identified in wheat that can cause severe immune diseases [204]. CD is caused by the ingestion of gluten proteins from wheat, leading to an autoimmune disorder [205]. In bread wheat, the downregulation of gliadins (the main toxic component of gluten) has been successfully achieved through RNAi technology [206,207]. In durum wheat, the 33-mer, the main immunodominant peptide of four highly stimulatory peptides encoding by the α-gliadin genes, was precisely modified through the CRISPR/Cas9 technology [187]. Compared to the wild type, the produced mutant lines showed a highest 69% reduction of gliadin in the seeds, displaying a reduced immunoreactivity [187]. On the other hand, the α-amylase/trypsin inhibitors (ATI) are the major triggering factors responsible for the onset of bakers’ asthma (develops after allergen inhalation) [208,209]. An ATI mutant durum wheat obtained through CRISPR/Cas9 multiplex editing knocked out two major related subunits, WTAI-CM3 and WTAI-CM16, showed reduced allergen proteins [188]. Compared to traditional breeding strategies, novel technologies focused on manipulating molecule levels tend to be more time-saving and effective, especially in modifying traits controlled by multiple genes and purging deleterious alleles.

4. Future Perspectives

Considering the substantial role that durum wheat has been playing in the global food supply, its yield and product diversity are related to a sustainable global development, especially upon facing severe climate change and rapidly growing populations. However, the long history of domestication and modern breeding programs have caused serious genetic erosion in durum wheat, leading to a bottleneck in yield improvement and vulnerability to biotic and abiotic stress. To broaden the genetic diversity of durum wheat, massive germplasm resources and approaches have been employed in the attempts of re-evolution of durum wheat as discussed above. Future improvement efforts should be focused on the following aspects.
First, durum wheat is a polyploid species that has experienced a long evolution. The artificial and natural selection caused huge variations between durum wheat and its subgenome donors. Species including T. urartu, T. boeoticum, T. monococcum, and Ae. speltoides are essential resources for the improvement of quality traits, resistance against several wheat diseases, and environmental adaptability. They could be used to transfer alien genes into durum wheat through inter-ploidy hybridization. Moreover, desirable genes or traits can also be introduced into durum wheat via ‘bridge-crossing’ using synthesized amphiploids, for instances AASS, AABBAmAm, and AABBSS [168,170,210,211,212,213].
Secondly, the tetraploid relatives of durum wheat and hexaploid wheat are critical germplasm for durum wheat breeding. For example, wild emmer wheat possesses many important beneficial traits, such as resistance to stripe rust, stem rust, and powdery mildew, and high tillering capacity, grain protein content, photosynthetic yield, and salt and drought tolerance [126]. The other cultivated turgidum wheat, i.e., Rivet wheat, Polish wheat, and Persian wheat, also contain useful genes [185]. The elite genes or traits carried by the ABD genomes of hexaploid wheat can be easily introgressed into durum wheat through crossing. In addition, the AAGG group wheat should also be explored for genetic enrichment of durum wheat [214].
Third, the availability of the whole genome of allohexaploid wheat CS and several allotetraploid wheat has eased the exploration of new genes in durum breeding. So far, numerous genes still have untapped functions. Conventional breeding should be combined with modern molecular biological technologies, such as CRISPR/Cas9 and RNAi. Through the introduction of the TaWOX5 gene, the transformation efficiency of tetraploid Polish wheat has been dramatically enhanced, which makes gene-editing an ideal approach for durum wheat improvement [215]. Multiple desirable alleles can be pyramided to meet different breeding requirements [216]. On the other hand, deleterious genes or alleles need to be silenced or knocked out, such as the brittle rachis and tough glume carried by wild emmer wheat. The desirable candidate genes can be transferred into durum wheat without linkage drag of deleterious genes.

Author Contributions

Conceptualization: W.Y., W.M. and Y.Z.; Visualization: Q.L. and H.L.; Writing—original draft: F.Y. and J.Z.; Writing—review & editing: W.Y. and W.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the China Scholarship Council (Grant number: CSC201906910028).

Acknowledgments

We would like to thank Maoyun She, Shahidul Islam, and Nigarin Sultana for their help on the collection of literature.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sousa, D.T.; Ribeiro, M.; Sabença, C.; Igrejas, G. The 10,000-Year Success Story of Wheat! Foods 2021, 10, 2124. [Google Scholar] [CrossRef] [PubMed]
  2. Kenzhebayeva, S.; Abekova, A.; Atabayeva, S.; Yernazarova, G.; Omirbekova, N.; Zhang, G.; Turasheva, S.; Asrandina, S.; Sarsu, F.; Wang, Y. Mutant lines of spring wheat with increased iron, zinc, and micronutrients in grains and enhanced bioavailability for human health. BioMed Res. Int. 2019, 2019, 9692053. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Chen, X.-P.; Zhang, Y.-Q.; Tong, Y.-P.; Xue, Y.-F.; Liu, D.-Y.; Zhang, W.; Deng, Y.; Meng, Q.-F.; Yue, S.-C.; Yan, P. Harvesting more grain zinc of wheat for human health. Sci. Rep. 2017, 7, 7016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Shewry, P.R. Wheat. J. Exp. Bot. 2009, 60, 1537–1553. [Google Scholar] [CrossRef] [PubMed]
  5. Faltermaier, A.; Waters, D.; Becker, T.; Arendt, E.; Gastl, M. Common wheat (Triticum aestivum L.) and its use as a brewing cereal—A review. J. Inst. Brew. 2014, 120, 1–15. [Google Scholar] [CrossRef]
  6. Jouanin, A.; Gilissen, L.J.; Boyd, L.A.; Cockram, J.; Leigh, F.J.; Wallington, E.J.; Van den Broeck, H.C.; Van der Meer, I.M.; Schaart, J.G.; Visser, R.G. Food processing and breeding strategies for coeliac-safe and healthy wheat products. Food Res. Int. 2018, 110, 11–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Curtis, T.; Halford, N. Food security: The challenge of increasing wheat yield and the importance of not compromising food safety. Ann. Appl. Biol. 2014, 164, 354–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Tanaka, T.; Guo, J. How does the self-sufficiency rate affect international price volatility transmissions in the wheat sector? Evidence from wheat-exporting countries. Humanit. Soc. Sci. Commun. 2020, 7, 113. [Google Scholar] [CrossRef]
  9. Tripathi, A.D.; Mishra, R.; Maurya, K.K.; Singh, R.B.; Wilson, D.W. Estimates for world population and global food availability for global health. In The Role of Functional Food Security in Global Health; Elsevier: Amsterdam, The Netherlands, 2019; pp. 3–24. [Google Scholar]
  10. International Wheat Genome Sequencing Consortium (IWGSC). Wheat—A Key Crop for Food Security. Available online: http://www.wheatgenome.org/News/Media-resources/Fact-Sheets-Infographics/Wheat-A-Key-Crop-for-Food-Security (accessed on 18 April 2018).
  11. Enghiad, A.; Ufer, D.; Countryman, A.M.; Thilmany, D.D. An overview of global wheat market fundamentals in an era of climate concerns. Int. J. Agron. 2017, 2017, 3931897. [Google Scholar] [CrossRef]
  12. Bajwa, A.A.; Farooq, M.; Al-Sadi, A.M.; Nawaz, A.; Jabran, K.; Siddique, K.H. Impact of climate change on biology and management of wheat pests. Crop Protect. 2020, 137, 105304. [Google Scholar] [CrossRef]
  13. Harkness, C.; Semenov, M.A.; Areal, F.; Senapati, N.; Trnka, M.; Balek, J.; Bishop, J. Adverse weather conditions for UK wheat production under climate change. Agric. For. Meteorol. 2020, 282, 107862. [Google Scholar] [CrossRef] [PubMed]
  14. Hernandez-Ochoa, I.M.; Asseng, S.; Kassie, B.T.; Xiong, W.; Robertson, R.; Pequeno, D.N.L.; Sonder, K.; Reynolds, M.; Babar, M.A.; Milan, A.M. Climate change impact on Mexico wheat production. Agric. For. Meteorol. 2018, 263, 373–387. [Google Scholar] [CrossRef]
  15. Miedaner, T.; Juroszek, P. Climate change will influence disease resistance breeding in wheat in Northwestern Europe. Theor. Appl. Genet. 2021, 134, 1771–1785. [Google Scholar] [CrossRef] [PubMed]
  16. Xiao, D.; Li Liu, D.; Wang, B.; Feng, P.; Bai, H.; Tang, J. Climate change impact on yields and water use of wheat and maize in the North China Plain under future climate change scenarios. Agric. Water Manag. 2020, 238, 106238. [Google Scholar] [CrossRef]
  17. Yen, C.; Yang, J. Biosystematics of Triticeae: Volume I. Triticum-Aegilops Complex; Springer Nature: Cham, Switzerland, 2020. [Google Scholar]
  18. Kihara, H. Discovery of the DD–analyzer, one of the analyzers of Triticum vulgare. Agric. Hortic. 1944, 19, 889–890. [Google Scholar]
  19. McFadden, E.S.; Sears, E.R. The origin of Triticum spelta and its free-threshing hexaploid relatives. J. Hered. 1946, 37, 81–89. [Google Scholar] [CrossRef]
  20. Peng, J.; Sun, D.; Nevo, E. Wild emmer wheat, Triticum dicoccoides, occupies a pivotal position in wheat domestication process. Aust. J. Crop Sci. 2011, 5, 1127–1143. [Google Scholar]
  21. Dubcovsky, J.; Dvorak, J. Genome plasticity a key factor in the success of polyploid wheat under domestication. Science 2007, 316, 1862–1866. [Google Scholar] [CrossRef] [Green Version]
  22. Li, A.; Liu, D.; Yang, W.; Kishii, M.; Mao, L. Synthetic hexaploid wheat: Yesterday, today, and tomorrow. Engineering 2018, 4, 552–558. [Google Scholar] [CrossRef]
  23. Agriculture and Agri-Food Canada. Canada: Outlook for Principal Field Crops. Available online: https://www.agr.gc.ca/eng/canadas-agriculture-sectors/crops/reports-and-statistics-data-for-canadian-principal-field-crops/canada-outlook-for-principal-field-crops-2021-04-20/?id=1619114343630 (accessed on 20 April 2021).
  24. Tidiane Sall, A.; Chiari, T.; Legesse, W.; Seid-Ahmed, K.; Ortiz, R.; Van Ginkel, M.; Bassi, F.M. Durum wheat (Triticum durum Desf.): Origin, cultivation and potential expansion in Sub-Saharan Africa. Agronomy 2019, 9, 263. [Google Scholar] [CrossRef] [Green Version]
  25. Xynias, I.N.; Mylonas, I.; Korpetis, E.G.; Ninou, E.; Tsaballa, A.; Avdikos, I.D.; Mavromatis, A.G. Durum wheat breeding in the Mediterranean region: Current status and future prospects. Agronomy 2020, 10, 432. [Google Scholar] [CrossRef] [Green Version]
  26. Matsuoka, Y. Evolution of polyploid Triticum wheats under cultivation: The role of domestication, natural hybridization and allopolyploid speciation in their diversification. Plant Cell Physiol. 2011, 52, 750–764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Hao, M.; Zhang, L.; Ning, S.; Huang, L.; Yuan, Z.; Wu, B.; Yan, Z.; Dai, S.; Jiang, B.; Zheng, Y. The Resurgence of Introgression Breeding, as Exemplified in Wheat Improvement. Front. Plant Sci. 2020, 11, 252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Morris, C.F.; Kiszonas, A.M.; Murray, J.; Boehm, J.D.; Ibba, M.I.; Zhang, M.; Cai, X. Re-evolution of durum wheat by introducing the Hardness and Glu-D1 loci. Front. Sustain. Food Syst. 2019, 3, 103. [Google Scholar] [CrossRef]
  29. Harlan, J.R.; Zohary, D. Distribution of wild wheats and barley. Science 1966, 153, 1074–1080. [Google Scholar] [CrossRef]
  30. Abbo, S.; Lev-Yadun, S.; Gopher, A. Agricultural origins: Centers and noncenters; a Near Eastern reappraisal. Crit. Rev. Plant Sci. 2010, 29, 317–328. [Google Scholar] [CrossRef]
  31. Rahman, S.; Islam, S.; Yu, Z.; She, M.; Nevo, E.; Ma, W. Current progress in understanding and recovering the wheat genes lost in evolution and domestication. Int. J. Mol. Sci. 2020, 21, 5836. [Google Scholar] [CrossRef]
  32. Dvořák, J.; Terlizzi, P.d.; Zhang, H.-B.; Resta, P. The evolution of polyploid wheats: Identification of the A genome donor species. Genome 1993, 36, 21–31. [Google Scholar] [CrossRef]
  33. Ling, H.-Q.; Zhao, S.; Liu, D.; Wang, J.; Sun, H.; Zhang, C.; Fan, H.; Li, D.; Dong, L.; Tao, Y. Draft genome of the wheat A-genome progenitor Triticum urartu. Nature 2013, 496, 87–90. [Google Scholar] [CrossRef] [Green Version]
  34. Haider, N. The origin of the B-genome of bread wheat (Triticum aestivum L.). Russ. J. Genet. 2013, 49, 263–274. [Google Scholar] [CrossRef]
  35. Zohary, D.; Hopf, M. Domestication of Plants in the Old World; Oxford University Press: Oxford, UK, 2000. [Google Scholar]
  36. Johnson, B.; Dhaliwal, H. Reproductive isolation of Triticum boeoticum and Triticum urartu and the origin of the tetraploid wheats. Am. J. Bot. 1976, 63, 1088–1094. [Google Scholar] [CrossRef]
  37. Heun, M.; Schäfer-Pregl, R.; Klawan, D.; Castagna, R.; Accerbi, M.; Borghi, B.; Salamini, F. Site of einkorn wheat domestication identified by DNA fingerprinting. Science 1997, 278, 1312–1314. [Google Scholar] [CrossRef] [Green Version]
  38. Feldman, M.; Sears, E.R. The wild gene resources of wheat. Sci. Am. 1981, 244, 102–113. [Google Scholar] [CrossRef]
  39. Watanabe, N. Breeding opportunities for early, free-threshing and semi-dwarf Triticum monococcum L. Euphytica 2017, 213, 201. [Google Scholar] [CrossRef]
  40. Feldman, M.; Levy, A. Allopolyploidy—A shaping force in the evolution of wheat genomes. Cytogenet. Genome Res. 2005, 109, 250–258. [Google Scholar] [CrossRef]
  41. Pont, C.; Leroy, T.; Seidel, M.; Tondelli, A.; Duchemin, W.; Armisen, D.; Lang, D.; Bustos-Korts, D.; Goué, N.; Balfourier, F. Tracing the ancestry of modern bread wheats. Nat. Genet. 2019, 51, 905–911. [Google Scholar] [CrossRef]
  42. Daud, H.M.; Gustafson, J. Molecular evidence for Triticum speltoides as a B-genome progenitor of wheat (Triticum aestivum). Genome 1996, 39, 543–548. [Google Scholar] [CrossRef] [PubMed]
  43. Baum, B.; Edwards, T.; Johnson, D. Phylogenetic relationships among diploid Aegilops species inferred from 5S rDNA units. Mol. Phylogen. Evol. 2009, 53, 34–44. [Google Scholar] [CrossRef]
  44. Golovnina, K.; Glushkov, S.; Blinov, A.; Mayorov, V.; Adkison, L.; Goncharov, N. Molecular phylogeny of the genus Triticum L. Plant Syst. Evol. 2007, 264, 195–216. [Google Scholar] [CrossRef]
  45. Li, L.-F.; Zhang, Z.-B.; Wang, Z.-H.; Li, N.; Sha, Y.; Wang, X.-F.; Ding, N.; Li, Y.; Zhao, J.; Wu, Y.; et al. Genome sequences of the five Sitopsis species of Aegilops and the origin of polyploid wheat B-subgenome. Mol. Plant 2022. [Google Scholar] [CrossRef]
  46. Sears, E. The cytology and genetics of the wheats and their relatives. Adv. Genet. 1948, 2, 239–270. [Google Scholar]
  47. Huang, S.; Sirikhachornkit, A.; Su, X.; Faris, J.; Gill, B.; Haselkorn, R.; Gornicki, P. Genes encoding plastid acetyl-CoA carboxylase and 3-phosphoglycerate kinase of the Triticum/Aegilops complex and the evolutionary history of polyploid wheat. Proc. Natl. Acad. Sci. USA 2002, 99, 8133–8138. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Peng, J.H.; Sun, D.; Nevo, E. Domestication evolution, genetics and genomics in wheat. Mol. Breed. 2011, 28, 281–301. [Google Scholar] [CrossRef]
  49. Wagenaar, E. Studies on the genome constitution of Triticum timopheevi Zhuk. II. The T. timopheevi complex and its origin. Evolution 1966, 20, 150–164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Charmet, G. Wheat domestication: Lessons for the future. Comptes Rendus Biol. 2011, 334, 212–220. [Google Scholar] [CrossRef]
  51. Dvorak, J.; Akhunov, E.D. Tempos of gene locus deletions and duplications and their relationship to recombination rate during diploid and polyploid evolution in the Aegilops-Triticum alliance. Genetics 2005, 171, 323–332. [Google Scholar] [CrossRef] [Green Version]
  52. Feldman, M.; Kislev, M.E. Domestication of emmer wheat and evolution of free-threshing tetraploid wheat. Isr. J. Plant Sci. 2007, 55, 207–221. [Google Scholar] [CrossRef]
  53. Kislev, M.E.; Nadel, D.; Carmi, I. Epipalaeolithic (19,000 BP) cereal and fruit diet at Ohalo II, Sea of Galilee, Israel. Rev. Palaeobot. Palynol. 1992, 73, 161–166. [Google Scholar] [CrossRef]
  54. Özkan, H.; Willcox, G.; Graner, A.; Salamini, F.; Kilian, B. Geographic distribution and domestication of wild emmer wheat (Triticum dicoccoides). Genet. Resour. Crop Evol. 2010, 58, 11–53. [Google Scholar] [CrossRef]
  55. Aaronsohn, A.; Schweinfurth, G. Die auffindung des wilden emmers (Triticum dicoccum) in Nordpalästina. In Altneuland Monatsschrift für die Irtschaft Erschliessung Palästinas; Kessinger Publishing: Whitefish, MT, USA, 1906; Volume 7, pp. 213–220. [Google Scholar]
  56. Tanno, K.-I.; Willcox, G. How fast was wild wheat domesticated? Science 2006, 311, 1886. [Google Scholar] [CrossRef] [Green Version]
  57. Mori, N.; Moriguchi, T.; Nakamura, C. RFLP analysis of nuclear DNA for study of phylogeny and domestication of tetraploid wheat. Genes Genet. Syst. 1997, 72, 153–161. [Google Scholar] [CrossRef] [Green Version]
  58. Nesbitt, M. Wheat evolution: Integrating archaeological and biological evidence. In Wheat Taxonomy: The Legacy of John Percival; The Linnean Society of London: London, UK, 2001. [Google Scholar]
  59. Feldman, M. Origin of cultivated wheat. In The World Wheat Book; Bonjean, A.P., Angus, W.J., Eds.; Lavoisier Publishing: Paris, France, 2001; pp. 3–56. [Google Scholar]
  60. Salamini, F.; Özkan, H.; Brandolini, A.; Schäfer-Pregl, R.; Martin, W. Genetics and geography of wild cereal domestication in the near east. Nat. Rev. Genet. 2002, 3, 429–441. [Google Scholar] [CrossRef] [PubMed]
  61. Nave, M.; Taş, M.; Raupp, J.; Tiwari, V.K.; Ozkan, H.; Poland, J.; Hale, I.; Komatsuda, T.; Distelfeld, A. The independent domestication of Timopheev’s wheat: Insights from haplotype analysis of the Brittle rachis 1 (BTR1-A) gene. Genes 2021, 12, 338. [Google Scholar] [CrossRef] [PubMed]
  62. Mori, N.; Kondo, Y.; Ishii, T.; Kawahara, T.; Nakamura, C. Genetic diversity and origin of timopheevi wheat inferred by chloroplast DNA fingerprinting. Breed. Sci. 2009, 59, 571–578. [Google Scholar] [CrossRef] [Green Version]
  63. Luo, M.-C.; Yang, Z.-L.; You, F.; Kawahara, T.; Waines, J.; Dvorak, J. The structure of wild and domesticated emmer wheat populations, gene flow between them, and the site of emmer domestication. Theor. Appl. Genet. 2007, 114, 947–959. [Google Scholar] [CrossRef] [PubMed]
  64. Upadhya, M.; Swaminathan, M. Genome analysis in Triticum zhukovskyi, a new hexaploid wheat. Chromosoma 1963, 14, 589–600. [Google Scholar] [CrossRef]
  65. Mori, N.; Ishii, T.; Ishido, T.; Hirosawa, S.; Watatani, H.; Kawahara, T.; Nesbitt, M.; Belay, G.; Takumi, S.; Ogihara, Y.; et al. Origins of domesticated emmer and common wheat inferred from chloroplast DNA fingerprinting. In Proceedings of the Tenth International Wheat Genetics Symposium, Paestum, Italy, 1–6 September 2003; pp. 25–28. [Google Scholar]
  66. Salamini, F.; Heun, M.; Ozkan, H.; Wunder, J. On DNA markers, phylogenetic trees and mode of origin of crops. Genome 2004, 46, 448–453. [Google Scholar]
  67. Özkan, H.; Brandolini, A.; Schäfer-Pregl, R.; Salamini, F. AFLP analysis of a collection of tetraploid wheats indicates the origin of emmer and hard wheat domestication in southeast Turkey. Mol. Biol. Evol. 2002, 19, 1797–1801. [Google Scholar] [CrossRef] [Green Version]
  68. Salamini, F.; Heun, M.; Brandolini, A.; Ozkan, H.; Wunder, J. Comment on “AFLP data and the origins of domesticated crops”. Genome 2004, 47, 615–620. [Google Scholar] [CrossRef] [Green Version]
  69. Özkan, H.; Brandolini, A.; Pozzi, C.; Effgen, S.; Wunder, J.; Salamini, F. A reconsideration of the domestication geography of tetraploid wheats. Theor. Appl. Genet. 2005, 110, 1052–1060. [Google Scholar] [CrossRef] [Green Version]
  70. Li, W.; Gill, B.S. Multiple genetic pathways for seed shattering in the grasses. Funct. Integr. Genom. 2006, 6, 300–309. [Google Scholar] [CrossRef] [PubMed]
  71. Peng, J.; Ronin, Y.; Fahima, T.; Röder, M.S.; Li, Y.; Nevo, E.; Korol, A. Domestication quantitative trait loci in Triticum dicoccoides, the progenitor of wheat. Proc. Natl. Acad. Sci. USA 2003, 100, 2489–2494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Balter, M. Seeking agriculture’s ancient roots. Science 2007, 316, 1830–1835. [Google Scholar] [CrossRef] [PubMed]
  73. Watanabe, N.; Fujii, Y.; Kato, N.; Ban, T.; Martinek, P. Microsatellite mapping of the genes for brittle rachis on homoeologous group 3 chromosomes in tetraploid and hexaploid wheats. J. Appl. Genet. 2006, 47, 93–98. [Google Scholar] [CrossRef]
  74. Chen, Q.-F.; Yen, C.; Yang, J.-L. Chromosome location of the gene for brittle rachis in the Tibetan weedrace of common wheat. Genet. Resour. Crop Evol. 1998, 45, 407–410. [Google Scholar] [CrossRef]
  75. Watanabe, N.; Takesada, N.; Fujii, Y.; Martinek, P. Comparative mapping of genes for brittle rachis in Triticum and Aegilops. Czech J. Genet. Plant Breed 2005, 41, 39–44. [Google Scholar] [CrossRef] [Green Version]
  76. Nalam, V.J.; Vales, M.I.; Watson, C.J.; Kianian, S.F.; Riera-Lizarazu, O. Map-based analysis of genes affecting the brittle rachis character in tetraploid wheat (Triticum turgidum L.). Theor. Appl. Genet. 2006, 112, 373–381. [Google Scholar] [CrossRef]
  77. Pourkheirandish, M.; Hensel, G.; Kilian, B.; Senthil, N.; Chen, G.; Sameri, M.; Azhaguvel, P.; Sakuma, S.; Dhanagond, S.; Sharma, R. Evolution of the grain dispersal system in barley. Cell 2015, 162, 527–539. [Google Scholar] [CrossRef] [Green Version]
  78. Avni, R.; Nave, M.; Barad, O.; Baruch, K.; Twardziok, S.O.; Gundlach, H.; Hale, I.; Mascher, M.; Spannagl, M.; Wiebe, K. Wild emmer genome architecture and diversity elucidate wheat evolution and domestication. Science 2017, 357, 93–97. [Google Scholar] [CrossRef] [Green Version]
  79. Pourkheirandish, M.; Dai, F.; Sakuma, S.; Kanamori, H.; Distelfeld, A.; Willcox, G.; Kawahara, T.; Matsumoto, T.; Kilian, B.; Komatsuda, T. On the origin of the non-brittle rachis trait of domesticated einkorn wheat. Front. Plant Sci. 2018, 8, 2031. [Google Scholar] [CrossRef] [Green Version]
  80. Cao, W.; Scoles, G.J.; Hucl, P. The genetics of rachis fragility and glume tenacity in semi-wild wheat. Euphytica 1997, 94, 119–124. [Google Scholar] [CrossRef]
  81. Watanabe, N.; Sugiyama, K.; Yamagishi, Y.; Sakata, Y. Comparative telosomic mapping of homoeologous genes for brittle rachis in tetraploid and hexaploid wheats. Hereditas 2002, 137, 180–185. [Google Scholar] [CrossRef]
  82. Peng, J.; Liu, Z.; Liu, X.; Yan, J.; Sun, D.; Nevo, E. Evolutionary agriculture domestication of wild emmer wheat. In New Horizons in Evolution; Elsevier: Amsterdam, The Netherlands, 2021; pp. 193–255. [Google Scholar]
  83. Sharma, J.S.; Running, K.L.; Xu, S.S.; Zhang, Q.; Haugrud, A.R.P.; Sharma, S.; McClean, P.E.; Faris, J.D. Genetic analysis of threshability and other spike traits in the evolution of cultivated emmer to fully domesticated durum wheat. Mol. Genet. Genom. 2019, 294, 757–771. [Google Scholar] [CrossRef] [PubMed]
  84. Dvorak, J.; Deal, K.R.; Luo, M.-C.; You, F.M.; von Borstel, K.; Dehghani, H. The origin of spelt and free-threshing hexaploid wheat. J. Hered. 2012, 103, 426–441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Kerber, E.; Rowland, G. Origin of the free threshing character in hexaploid wheat. Can. J. Genet. Cytol. 1974, 16, 145–154. [Google Scholar] [CrossRef]
  86. Chen, Q.-F.; Yen, C.; Yang, J.-L. Chromosome location of the gene for the hulled character in the Tibetan weedrace of common wheat. Genet. Resour. Crop Evol. 1999, 46, 543–546. [Google Scholar] [CrossRef]
  87. Nalam, V.J.; Vales, M.I.; Watson, C.J.; Johnson, E.B.; Riera-Lizarazu, O. Map-based analysis of genetic loci on chromosome 2D that affect glume tenacity and threshability, components of the free-threshing habit in common wheat (Triticum aestivum L.). Theor. Appl. Genet. 2007, 116, 135–145. [Google Scholar] [CrossRef]
  88. Faris, J.D.; Gill, B.S. Genomic targeting and high-resolution mapping of the domestication gene Q in wheat. Genome 2002, 45, 706–718. [Google Scholar] [CrossRef]
  89. Simons, K.J.; Fellers, J.P.; Trick, H.N.; Zhang, Z.; Tai, Y.-S.; Gill, B.S.; Faris, J.D. Molecular characterization of the major wheat domestication gene Q. Genetics 2006, 172, 547–555. [Google Scholar] [CrossRef] [Green Version]
  90. Song, G.; Sun, G.; Kong, X.; Jia, M.; Wang, K.; Ye, X.; Zhou, Y.; Geng, S.; Mao, L.; Li, A. The soft glumes of common wheat are sterile-lemmas as determined by the domestication gene Q. Crop J. 2019, 7, 113–117. [Google Scholar] [CrossRef]
  91. Faris, J.D.; Zhang, Z.; Chao, S. Map-based analysis of the tenacious glume gene Tg-B1 of wild emmer and its role in wheat domestication. Gene 2014, 542, 198–208. [Google Scholar] [CrossRef] [PubMed]
  92. Faris, J.D.; Zhang, Q.; Chao, S.; Zhang, Z.; Xu, S.S. Analysis of agronomic and domestication traits in a durum × cultivated emmer wheat population using a high-density single nucleotide polymorphism-based linkage map. Theor. Appl. Genet. 2014, 127, 2333–2348. [Google Scholar] [CrossRef] [PubMed]
  93. Jantasuriyarat, C.; Vales, M.; Watson, C.; Riera-Lizarazu, O. Identification and mapping of genetic loci affecting the free-threshing habit and spike compactness in wheat (Triticum aestivum L.). Theor. Appl. Genet. 2004, 108, 261–273. [Google Scholar] [CrossRef] [PubMed]
  94. Liu, H.; Wang, K.; Tang, H.; Gong, Q.; Du, L.; Pei, X.; Ye, X. CRISPR/Cas9 editing of wheat TaQ genes alters spike morphogenesis and grain threshability. J. Genet. Genom. 2020, 47, 563–575. [Google Scholar] [CrossRef] [PubMed]
  95. Debernardi, J.M.; Lin, H.; Chuck, G.; Faris, J.D.; Dubcovsky, J. microRNA172 plays a crucial role in wheat spike morphogenesis and grain threshability. Development 2017, 144, 1966–1975. [Google Scholar] [PubMed] [Green Version]
  96. Greenwood, J.R.; Finnegan, E.J.; Watanabe, N.; Trevaskis, B.; Swain, S.M. New alleles of the wheat domestication gene Q reveal multiple roles in growth and reproductive development. Development 2017, 144, 1959–1965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Zhang, Z.; Belcram, H.; Gornicki, P.; Charles, M.; Just, J.; Huneau, C.; Magdelenat, G.; Couloux, A.; Samain, S.; Gill, B.S.; et al. Duplication and partitioning in evolution and function of homoeologous Q loci governing domestication characters in polyploid wheat. Proc. Natl. Acad. Sci. USA 2011, 108, 18737–18742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Muramatsu, M. Dosage effect of the spelta gene q of hexaploid wheat. Genetics 1963, 48, 469. [Google Scholar] [CrossRef] [PubMed]
  99. Faris, J.; Simons, K.J.; Zhang, Z.; Gill, B. The wheat super domestication gene Q. Front. Wheat Biosci. 2005, 129–148. [Google Scholar]
  100. Vaughan, D.; Balazs, E.; Heslop-Harrison, J. From crop domestication to super-domestication. Ann. Bot. 2007, 100, 893–901. [Google Scholar] [CrossRef]
  101. Xie, Q.; Li, N.; Yang, Y.; Lv, Y.; Yao, H.; Wei, R.; Sparkes, D.L.; Ma, Z. Pleiotropic effects of the wheat domestication gene Q on yield and grain morphology. Planta 2018, 247, 1089–1098. [Google Scholar] [CrossRef] [PubMed]
  102. Gegas, V.C.; Nazari, A.; Griffiths, S.; Simmonds, J.; Fish, L.; Orford, S.; Sayers, L.; Doonan, J.H.; Snape, J.W. A genetic framework for grain size and shape variation in wheat. Plant Cell 2010, 22, 1046–1056. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Steiger, D.; Elias, E.; Cantrell, R. Evaluation of lines derived from wild emmer chromosome substitutions: I. Quality traits. Crop Sci. 1996, 36, 223–227. [Google Scholar] [CrossRef]
  104. Campbell, B.; Baenziger, P.S.; Gill, K.; Eskridge, K.M.; Budak, H.; Erayman, M.; Dweikat, I.M.; Yen, Y. Identification of QTLs and environmental interactions associated with agronomic traits on chromosome 3A of wheat. Crop Sci. 2003, 43, 1493–1505. [Google Scholar] [CrossRef]
  105. Kato, K.; Miura, H.; Sawada, S. Mapping QTLs controlling grain yield and its components on chromosome 5A of wheat. Theor. Appl. Genet. 2000, 101, 1114–1121. [Google Scholar] [CrossRef]
  106. Liu, J.; Tang, H.; Qu, X.; Liu, H.; Li, C.; Tu, Y.; Li, S.; Habib, A.; Mu, Y.; Dai, S. A novel, major, and validated QTL for the effective tiller number located on chromosome arm 1BL in bread wheat. Plant Mol. Biol. 2020, 104, 173–185. [Google Scholar] [CrossRef]
  107. Spielmeyer, W.; Richards, R. Comparative mapping of wheat chromosome 1AS which contains the tiller inhibition gene (tin) with rice chromosome 5S. Theor. Appl. Genet. 2004, 109, 1303–1310. [Google Scholar] [CrossRef]
  108. Cockram, J.; Jones, H.; Leigh, F.J.; O’Sullivan, D.; Powell, W.; Laurie, D.A.; Greenland, A.J. Control of flowering time in temperate cereals: Genes, domestication, and sustainable productivity. J. Exp. Bot. 2007, 58, 1231–1244. [Google Scholar] [CrossRef]
  109. Würschum, T.; Langer, S.M.; Longin, C.F.H.; Tucker, M.R.; Leiser, W.L. A modern Green Revolution gene for reduced height in wheat. Plant J. 2017, 92, 892–903. [Google Scholar] [CrossRef] [Green Version]
  110. Pearce, S.; Saville, R.; Vaughan, S.P.; Chandler, P.M.; Wilhelm, E.P.; Sparks, C.A.; Al-Kaff, N.; Korolev, A.; Boulton, M.I.; Phillips, A.L. Molecular characterization of Rht-1 dwarfing genes in hexaploid wheat. Plant Physiol. 2011, 157, 1820–1831. [Google Scholar] [CrossRef] [Green Version]
  111. Hedden, P. The genes of the Green Revolution. Trends Genet. 2003, 19, 5–9. [Google Scholar] [CrossRef]
  112. Chai, S.; Yao, Q.; Zhang, X.; Xiao, X.; Fan, X.; Zeng, J.; Sha, L.; Kang, H.; Zhang, H.; Li, J. The semi-dwarfing gene Rht-dp from dwarf polish wheat (Triticum polonicum L.) is the “Green Revolution” gene Rht-B1b. BMC Genom. 2021, 22, 1–15. [Google Scholar] [CrossRef] [PubMed]
  113. Kang, H.-Y.; Lin, L.-J.; Song, Z.-J.; Yuan, J.-Y.; Zhong, M.-Y.; Zhang, H.-Q.; Fan, X.; Sha, L.-N.; Wang, Y.; Xu, L.-L. Identification, fine mapping and characterization of Rht-dp, a recessive wheat dwarfing (reduced height) gene derived from Triticum polonicum. Genes Genom. 2012, 34, 509–515. [Google Scholar] [CrossRef]
  114. Zaharieva, M.; Monneveux, P. Cultivated einkorn wheat (Triticum monococcum L. subsp. monococcum): The long life of a founder crop of agriculture. Genet. Resour. Crop Evol. 2014, 61, 677–706. [Google Scholar] [CrossRef]
  115. Arzani, A. Emmer (Triticum turgidum ssp. dicoccum) flour and bread. In Flour and Breads and Their Fortification in Health and Disease Prevention; Elsevier: Amsterdam, The Netherlands, 2019; pp. 89–98. [Google Scholar]
  116. Cooper, R. Re-discovering ancient wheat varieties as functional foods. J. Tradit. Complementary Med. 2015, 5, 138–143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Magallanes-López, A.M.; Ammar, K.; Morales-Dorantes, A.; González-Santoyo, H.; Crossa, J.; Guzmán, C. Grain quality traits of commercial durum wheat varieties and their relationships with drought stress and glutenins composition. J. Cereal Sci. 2017, 75, 1–9. [Google Scholar] [CrossRef]
  118. Elias, E. Durum wheat products. In Durum Wheat Quality in the Mediterranean Region; Options Méditerranéennes: Série A. Séminaires Méditerranéens, n. 22; International Center for Advanced Mediterranean Agronomic Studies (CIHEAM): Saragosa, Spain, 1995; Volume 40, pp. 23–31. [Google Scholar]
  119. International Pasta Organisation (IPO). Pasta, World Consumption Boom Since The Lockdown Started 1 Consumer Out of 4 Ate More, 25% Export Increase in 6 Months. Available online: https://internationalpasta.org/news/pasta-world-consumption-boom-since-the-lockdown-started-1-consumer-out-of-4-ate-more-25-export-increase-in-6-months/ (accessed on 16 October 2020).
  120. Maccaferri, M.; Harris, N.S.; Twardziok, S.O.; Pasam, R.K.; Gundlach, H.; Spannagl, M.; Ormanbekova, D.; Lux, T.; Prade, V.M.; Milner, S.G. Durum wheat genome highlights past domestication signatures and future improvement targets. Nat. Genet. 2019, 51, 885–895. [Google Scholar] [CrossRef] [Green Version]
  121. Gustafson, P.; Raskina, O.; Ma, X.; Nevo, E. Wheat evolution, domestication, and improvement. In Making of a Wheat Plant; John Wiley & Sons (Wiley): Hoboken, NJ, USA, 2009; pp. 5–30. [Google Scholar]
  122. Padmanaban, S.; Zhang, P.; Hare, R.A.; Sutherland, M.W.; Martin, A. Pentaploid wheat hybrids: Applications, characterisation, and challenges. Front. Plant Sci. 2017, 8, 358. [Google Scholar] [CrossRef] [Green Version]
  123. Fu, Y.B.; Somers, D.J. Genome-wide reduction of genetic diversity in wheat breeding. Crop Sci. 2009, 49, 161–168. [Google Scholar] [CrossRef]
  124. Haudry, A.; Cenci, A.; Ravel, C.; Bataillon, T.; Brunel, D.; Poncet, C.; Hochu, I.; Poirier, S.; Santoni, S.; Glémin, S. Grinding up wheat: A massive loss of nucleotide diversity since domestication. Mol. Biol. Evol. 2007, 24, 1506–1517. [Google Scholar] [CrossRef] [Green Version]
  125. Nevo, E. Evolution of wild emmer wheat and crop improvement. J. Syst Evol. 2014, 52, 673–696. [Google Scholar] [CrossRef]
  126. Xie, W.; Nevo, E. Wild emmer: Genetic resources, gene mapping and potential for wheat improvement. Euphytica 2008, 164, 603–614. [Google Scholar] [CrossRef]
  127. El Haddad, N.; Kabbaj, H.; Zaïm, M.; El Hassouni, K.; Tidiane Sall, A.; Azouz, M.; Ortiz, R.; Baum, M.; Amri, A.; Gamba, F. Crop wild relatives in durum wheat breeding: Drift or thrift? Crop Sci. 2021, 61, 37–54. [Google Scholar] [CrossRef]
  128. Ogbonnaya, F.; Halloran, G.; Lagudah, E. D genome of wheat—60 years on from Kihara, Sears and McFadden. Front. Wheat BioSci. 2005, 11, 205–220. [Google Scholar]
  129. Mastrangelo, A.M.; Cattivelli, L. What Makes Bread and Durum Wheat Different? Trends Plant Sci. 2021, 26, 677–684. [Google Scholar] [CrossRef]
  130. Li, A.; Liu, D.; Wu, J.; Zhao, X.; Hao, M.; Geng, S.; Yan, J.; Jiang, X.; Zhang, L.; Wu, J. mRNA and small RNA transcriptomes reveal insights into dynamic homoeolog regulation of allopolyploid heterosis in nascent hexaploid wheat. Plant Cell 2014, 26, 1878–1900. [Google Scholar] [CrossRef] [Green Version]
  131. Sharma, H.; Waines, J.; Foster, K. Variability in Primitive and Wild Wheats for Useful Genetic Characters 1. Crop Sci. 1981, 21, 555–559. [Google Scholar] [CrossRef]
  132. Monneveux, P.; Zaharieva, M.; Rekika, D.; Royo, C.; Nachit, M.; Fonzo, N.; Araus, J. The utilisation of Triticum and Aegilops species for the improvement of durum wheat. In Durum Wheat Improvement in the Mediterranean Region: New Challenges; Options Méditerranéennes: Série A. Séminaires Méditerranéens, n. 40; International Center for Advanced Mediterranean Agronomic Studies (CIHEAM): Saragosa, Spain, 2000; pp. 71–81. [Google Scholar]
  133. Mohammadi, M.; Mirlohi, A.; Majidi, M.M.; Kartalaei, E.S. Emmer wheat as a source for trait improvement in durum wheat: A study of general and specific combining ability. Euphytica 2021, 217, 64. [Google Scholar] [CrossRef]
  134. Giorgi, B. A homoeologous pairing mutant isolated in Triticum durum cv. Cappelli. Mutat. Breed. Newsl. 1978, 11, 4–5. [Google Scholar]
  135. Sears, E.R. Genetics society of canada award of excellence lecture an induced mutant with homoeologous pairing in common wheat. Can. J. Genet. Cytol. 1977, 19, 585–593. [Google Scholar] [CrossRef]
  136. Joppa, L.; Williams, N. Langdon durum disomic substitution lines and aneuploid analysis in tetraploid wheat. Genome 1988, 30, 222–228. [Google Scholar] [CrossRef]
  137. Joppa, L.; Williams, N. A Disomic-5D-Nullisomic-5B Substitution Line of Durum Wheat 1. Crop Sci. 1979, 19, 509–511. [Google Scholar] [CrossRef]
  138. Joppa, L. Development of D-genome Disomic Substitution lines of Durum wheat (Triticum turgidum L.). In Fourth International Wheat Genetics Symposium Cytogenetics; University of Missouri-Columbia: Columbia, MI, USA, 1973; pp. 685–690. [Google Scholar]
  139. Gupta, P.; Langridge, P.; Mir, R. Marker-assisted wheat breeding: Present status and future possibilities. Mol. Breed. 2010, 26, 145–161. [Google Scholar] [CrossRef]
  140. Varshney, R.K.; Bohra, A.; Yu, J.; Graner, A.; Zhang, Q.; Sorrells, M.E. Designing future crops: Genomics-assisted breeding comes of age. Trends Plant Sci. 2021, 26, 631–649. [Google Scholar] [CrossRef] [PubMed]
  141. Hickey, L.T.; Hafeez, A.N.; Robinson, H.; Jackson, S.A.; Leal-Bertioli, S.C.; Tester, M.; Gao, C.; Godwin, I.D.; Hayes, B.J.; Wulff, B.B. Breeding crops to feed 10 billion. Nat. Biotechnol. 2019, 37, 744–754. [Google Scholar] [CrossRef] [Green Version]
  142. Liu, Q.; Yang, F.; Zhang, J.; Liu, H.; Rahman, S.; Islam, S.; Ma, W.; She, M. Application of CRISPR/Cas9 in Crop Quality Improvement. Int. J. Mol. Sci. 2021, 22, 4206. [Google Scholar] [CrossRef]
  143. Habash, D.; Kehel, Z.; Nachit, M. Genomic approaches for designing durum wheat ready for climate change with a focus on drought. J. Exp. Bot. 2009, 60, 2805–2815. [Google Scholar] [CrossRef]
  144. Siddiqui, H.A.; Harvey-Samuel, T.; Mansoor, S. Gene drive: A faster route to plant improvement. Trends Plant Sci. 2021, 26, 1204–1206. [Google Scholar] [CrossRef]
  145. McClung, A.M.; Cantrell, R.; Quick, J.; Gregory, R. Influence of the Rht1 Semidwarf Gene on Yield, Yield Components, and Grain Protein in Durum Wheat 1. Crop Sci. 1986, 26, 1095–1099. [Google Scholar] [CrossRef]
  146. Duan, S.; Zhao, Z.; Qiao, Y.; Cui, C.; Morgunov, A.; Condon, A.G.; Chen, L.; Hu, Y.-G. GAR dwarf gene Rht14 reduced plant height and affected agronomic traits in durum wheat (Triticum durum). Field Crop Res. 2020, 248, 107721. [Google Scholar] [CrossRef]
  147. Zhao, Z.; Duan, S.; Hao, J.; Cui, C.; Yang, Y.; Condon, A.G.; Chen, F.; Hu, Y.-G.; Chen, L. The dwarf gene Rht15 improved lodging resistance but differentially affected agronomic and quality traits in durum wheat. Field Crop Res. 2021, 263, 108058. [Google Scholar] [CrossRef]
  148. Ma, J.; Zhang, C.; Yan, G.; Liu, C. Improving yield and quality traits of durum wheat by introgressing chromosome segments from hexaploid wheat. Genet. Mol. Res. 2013, 12, 6120–6129. [Google Scholar] [CrossRef] [PubMed]
  149. Yang, F.; Liu, Q.; Wang, Q.; Yang, N.; Li, J.; Wan, H.; Liu, Z.; Yang, S.; Wang, Y.; Zhang, J. Characterization of the Durum Wheat-Aegilops tauschii 4D (4B) Disomic Substitution Line YL-443 With Superior Characteristics of High Yielding and Stripe Rust Resistance. Front. Plant Sci. 2021, 12, 745290. [Google Scholar] [CrossRef] [PubMed]
  150. Sestili, F.; Pagliarello, R.; Zega, A.; Saletti, R.; Pucci, A.; Botticella, E.; Masci, S.; Tundo, S.; Moscetti, I.; Foti, S. Enhancing grain size in durum wheat using RNAi to knockdown GW2 genes. Theor. Appl. Genet. 2019, 132, 419–429. [Google Scholar] [CrossRef] [PubMed]
  151. Gazza, L.; Zanella, L.; Pogna, N. Development of durum wheat (Triticum turgidum ssp. durum) lines with soft kernel texture by chromosome engineering. In Proceedings of the 11th International Wheat Genetics Symposium, Brisbane, Australia, 24–29 August 2008; Sydney University Press: Sydney, Australia, 2008. [Google Scholar]
  152. Morris, C.F.; Simeone, M.C.; King, G.; Lafiandra, D. Transfer of soft kernel texture from Triticum aestivum to durum wheat, Triticum turgidum ssp. durum. Crop Sci. 2011, 51, 114–122. [Google Scholar] [CrossRef]
  153. Li, Y.; Wang, Q.; Li, X.; Xiao, X.; Sun, F.; Wang, C.; Hu, W.; Feng, Z.; Chang, J.; Chen, M. Coexpression of the high molecular weight glutenin subunit 1Ax1 and puroindoline improves dough mixing properties in durum wheat (Triticum turgidum L. ssp. durum). PLoS ONE 2012, 7, e50057. [Google Scholar]
  154. Sissons, M.; Pleming, D.; Sestili, F.; Lafiandra, D. Effect of Glu-D1 gene introgression and amylose content on breadmaking potential of blends of durum and hexaploid wheat. Cereal Chem. 2019, 96, 193–206. [Google Scholar] [CrossRef]
  155. Ceoloni, C.; Ciaffi, M.; Lafiandra, D.; Giorgi, B. Chromosome engineering as a means of transferring 1D storage protein genes from common to durum wheat. In Proceedings of the Eighth International Wheat Genetics Symposium, Beijing, China, 20–25 July 1993; pp. 159–163. [Google Scholar]
  156. Ceoloni, C.; Biagetti, M.; Ciaffi, M.; Forte, P.; Pasquini, M. Wheat chromosome engineering at the 4x level: The potential of different alien gene transfers into durum wheat. Euphytica 1996, 89, 87–97. [Google Scholar] [CrossRef]
  157. Vitellozzi, F.; Ciaffi, M.; Dominici, L.; Ceoloni, C. Isolation of a chromosomally engineered durum wheat line carrying the common wheat Glu-D1d allele. Agronomie 1997, 17, 413–419. [Google Scholar] [CrossRef] [Green Version]
  158. Gennaro, A.; Forte, P.; Panichi, D.; Lafiandra, D.; Pagnotta, M.A.; D’Egidio, M.G.; Ceoloni, C. Stacking small segments of the 1D chromosome of bread wheat containing major gluten quality genes into durum wheat: Transfer strategy and breeding prospects. Mol. Breed. 2012, 30, 149–167. [Google Scholar] [CrossRef]
  159. Lukaszewski, A. Registration of six germplasms of bread wheat having variations of cytogenetically engineered wheat-rye translocation 1RS. 1BL.(Registrations Of Germplasm). Crop Sci. 2003, 43, 1137–1139. [Google Scholar] [CrossRef]
  160. Kovacs, M.; Howes, N.; Clarke, J.; Leisle, D. Quality Characteristics of Durum Wheat Lines Deriving High Protein From a Triticum dicoccoides (6b) Substitution. J. Cereal Sci. 1998, 27, 47–51. [Google Scholar] [CrossRef]
  161. Yu, S.; Li, M.; Dubcovsky, J.; Tian, L. Mutant combinations of lycopene ε-cyclase and β-carotene hydroxylase 2 homoeologs increased β-carotene accumulation in endosperm of tetraploid wheat (Triticum turgidum L.) grains. Plant Biotechnol. J. 2021, 11, 113. [Google Scholar] [CrossRef] [PubMed]
  162. Liu, Y.; Liu, Y.; Zhou, Y.; Wight, C.; Pu, Z.; Qi, P.; Jiang, Q.; Deng, M.; Wang, Z.; Wei, Y. Conferring resistance to pre-harvest sprouting in durum wheat by a QTL identified in Triticum spelta. Euphytica 2017, 213, 19. [Google Scholar] [CrossRef]
  163. Ceoloni, C.; Forte, P.; Gennaro, A.; Micali, S.; Carozza, R.; Bitti, A. Recent developments in durum wheat chromosome engineering. Cytogenet. Genome Res. 2005, 109, 328–334. [Google Scholar] [CrossRef]
  164. Ceoloni, C.; Kuzmanović, L.; Ruggeri, R.; Rossini, F.; Forte, P.; Cuccurullo, A.; Bitti, A. Harnessing genetic diversity of wild gene pools to enhance wheat crop production and sustainability: Challenges and opportunities. Diversity 2017, 9, 55. [Google Scholar] [CrossRef] [Green Version]
  165. Zaim, M.; El Hassouni, K.; Gamba, F.; Filali-Maltouf, A.; Belkadi, B.; Sourour, A.; Amri, A.; Nachit, M.; Taghouti, M.; Bassi, F.M. Wide crosses of durum wheat (Triticum durum Desf.) reveal good disease resistance, yield stability, and industrial quality across Mediterranean sites. Field Crop Res. 2017, 214, 219–227. [Google Scholar] [CrossRef]
  166. Rinaldo, A.; Gilbert, B.; Boni, R.; Krattinger, S.G.; Singh, D.; Park, R.F.; Lagudah, E.; Ayliffe, M. The Lr34 adult plant rust resistance gene provides seedling resistance in durum wheat without senescence. Plant Biotechnol. J. 2017, 15, 894–905. [Google Scholar] [CrossRef] [Green Version]
  167. Kerber, E.; Dyck, P. Inheritance of stem rust resistance transferred from diploid wheat (Triticum monococcum) to tetraploid and hexaploid wheat and chromosome location of the gene involved. Can. J. Genet. Cytol. 1973, 15, 397–409. [Google Scholar] [CrossRef]
  168. Ma, H.; Singh, R.; Mujeeb-Kazi, A. Resistance to stripe rust in durum wheats, A-genome diploids, and their amphiploids. Euphytica 1997, 94, 279–286. [Google Scholar] [CrossRef]
  169. Plamenov, D.; Belchev, I.; Kiryakova, V.; Spetsov, P. Fungal resistance of Triticum durum-T. monococcum ssp. aegilopoides amphiploid. J. Plant Dis. Prot. 2009, 116, 60–62. [Google Scholar] [CrossRef]
  170. Megyeri, M.; Mikó, P.; Molnár, I.; Kovács, G. Development of synthetic amphiploids based on different Triticum turgidum × T. monococcum crosses to improve the adaptability of cereals. Acta Agron. Hung. 2015, 59, 267–274. [Google Scholar] [CrossRef]
  171. Prat, N.; Guilbert, C.; Prah, U.; Wachter, E.; Steiner, B.; Langin, T.; Robert, O.; Buerstmayr, H. QTL mapping of Fusarium head blight resistance in three related durum wheat populations. Theor. Appl. Genet. 2017, 130, 13–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Zhao, M.; Leng, Y.; Chao, S.; Xu, S.S.; Zhong, S. Molecular mapping of QTL for Fusarium head blight resistance introgressed into durum wheat. Theor. Appl. Genet. 2018, 131, 1939–1951. [Google Scholar] [CrossRef] [PubMed]
  173. Giancaspro, A.; Giove, S.L.; Zito, D.; Blanco, A.; Gadaleta, A. Mapping QTLs for Fusarium head blight resistance in an interspecific wheat population. Front. Plant Sci. 2016, 7, 1381. [Google Scholar] [CrossRef] [PubMed]
  174. Giancaspro, A.; Lionetti, V.; Giove, S.L.; Zito, D.; Fabri, E.; Reem, N.; Zabotina, O.A.; De Angelis, E.; Monaci, L.; Bellincampi, D. Cell wall features transferred from common into durum wheat to improve Fusarium Head Blight resistance. Plant Sci. 2018, 274, 121–128. [Google Scholar] [CrossRef] [PubMed]
  175. Kumar, J.; Rai, K.M.; Pirseyedi, S.; Elias, E.M.; Xu, S.; Dill-Macky, R.; Kianian, S.F. Epigenetic regulation of gene expression improves Fusarium head blight resistance in durum wheat. Sci. Rep. 2020, 10, 17610. [Google Scholar] [CrossRef] [PubMed]
  176. Siham, O.; Abderraouf, H.; Saadia, L.; Nasserlhaq, N. Phenotypic characterization of recombinant inbred durum wheat lines segregating for resistance to Hessian fly (Mayetiola destructor (Say),) in Morocco. Rev. Nat. Technol. 2017, 9, 18. [Google Scholar]
  177. Rampino, P.; Mita, G.; Pataleo, S.; De Pascali, M.; Di Fonzo, N.; Perrotta, C. Acquisition of thermotolerance and HSP gene expression in durum wheat (Triticum durum Desf.) cultivars. Environ. Exp. Bot. 2009, 66, 257–264. [Google Scholar] [CrossRef]
  178. Khateeb, W.A.; Muhaidat, R.; Alahmed, S.; Al Zoubi, M.S.; Al-Batayneh, K.M.; El-Oqlah, A.; Gamar, M.A.; Hussein, E.; Aljabali, A.A.; Alkaraki, A.K. Heat shock proteins gene expression and physiological responses in durum wheat (Triticum durum) under salt stress. Physiol. Mol. Biol. Plants 2020, 26, 1599–1608. [Google Scholar] [CrossRef]
  179. Bacher, H.; Zhu, F.; Gao, T.; Liu, K.; Dhatt, B.K.; Awada, T.; Zhang, C.; Distelfeld, A.; Yu, H.; Peleg, Z. Wild emmer introgression alters root-to-shoot growth dynamics in durum wheat in response to water stress. Plant Physiol. 2021, 187, 1149–1162. [Google Scholar] [CrossRef] [PubMed]
  180. Ayadi, M.; Brini, F.; Masmoudi, K. Overexpression of a wheat aquaporin gene, TdPIP2; 1, enhances salt and drought tolerance in transgenic durum wheat cv. Maali. Int. J. Mol. Sci. 2019, 20, 2389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Azzedine, F.; Gherroucha, H.; Baka, M. Improvement of salt tolerance in durum wheat by ascorbic acid application. J. Stress Physiol. Biochem 2011, 7, 27–37. [Google Scholar]
  182. Luo, M.; Dubcovsky, J.; Goyal, S.; Dvořák, J. Engineering of interstitial foreign chromosome segments containing the K+/Na+ selectivity gene Kna1 by sequential homoeologous recombination in durum wheat. Theor. Appl. Genet. 1996, 93, 1180–1184. [Google Scholar] [CrossRef] [PubMed]
  183. Han, C.; Ryan, P.R.; Yan, Z.; Delhaize, E. Introgression of a 4D chromosomal fragment into durum wheat confers aluminium tolerance. Ann. Bot. 2014, 114, 135–144. [Google Scholar] [CrossRef] [Green Version]
  184. Han, C.; Zhang, P.; Ryan, P.R.; Rathjen, T.M.; Yan, Z.; Delhaize, E. Introgression of genes from bread wheat enhances the aluminium tolerance of durum wheat. Theor. Appl. Genet. 2016, 129, 729–739. [Google Scholar] [CrossRef]
  185. Munns, R.; Hare, R.; James, R.; Rebetzke, G. Genetic variation for improving the salt tolerance of durum wheat. Aust. J. Agric. Res. 2000, 51, 69–74. [Google Scholar] [CrossRef]
  186. Mohammadi, M.; Mirlohi, A.; Majidi, M.M.; Rabbani, A. Exploring the breeding potential of Iranian emmer wheats to increase durum wheat tolerance to drought. Plant Genet. Resour. 2021, 19, 363–374. [Google Scholar] [CrossRef]
  187. Sánchez-León, S.; Gil-Humanes, J.; Ozuna, C.V.; Giménez, M.J.; Sousa, C.; Voytas, D.F.; Barro, F. Low-gluten, nontransgenic wheat engineered with CRISPR/Cas9. Plant Biotechnol. J. 2018, 16, 902–910. [Google Scholar] [CrossRef]
  188. Camerlengo, F.; Frittelli, A.; Sparks, C.; Doherty, A.; Martignago, D.; Larré, C.; Lupi, R.; Sestili, F.; Masci, S. CRISPR-Cas9 multiplex editing of the α-amylase/trypsin inhibitor genes to reduce allergen proteins in durum wheat. Front. Sustain. Food Syst. 2020, 4, 104. [Google Scholar] [CrossRef]
  189. Velde, K.V.D.; Thomas, S.G.; Heyse, F.; Kaspar, R.; Straeten, D.V.D.; Rohde, A. N-terminal truncated RHT-1 proteins generated by translational reinitiation cause semi-dwarfing of wheat Green Revolution alleles. Mol. Plant 2021, 14, 679–687. [Google Scholar] [CrossRef] [PubMed]
  190. Mefleh, M.; Conte, P.; Fadda, C.; Giunta, F.; Piga, A.; Hassoun, G.; Motzo, R. From ancient to old and modern durum wheat varieties: Interaction among cultivar traits, management, and technological quality. J. Sci. Food Agric. 2019, 99, 2059–2067. [Google Scholar] [CrossRef] [PubMed]
  191. Morris, C.F. Puroindolines: The molecular genetic basis of wheat grain hardness. Plant Mol. Biol. 2002, 48, 633–647. [Google Scholar] [CrossRef] [PubMed]
  192. Bhave, M.; Morris, C.F. Molecular genetics of puroindolines and related genes: Allelic diversity in wheat and other grasses. Plant Mol. Biol. 2008, 66, 205–219. [Google Scholar] [CrossRef]
  193. Li, W.; Huang, L.; Gill, B.S. Recurrent Deletions of Puroindoline Genes at the Grain Hardness Locus in Four Independent Lineages of Polyploid Wheat. Plant Physiol. 2007, 146, 200–212. [Google Scholar] [CrossRef] [Green Version]
  194. Morris, C.F. Bread-baking quality and the effects of Glu-D1 gene introgressions in durum wheat (Triticum turgidum ssp. durum). Cereal Chem. 2021, 98, 1151–1158. [Google Scholar] [CrossRef]
  195. Colasuonno, P.; Marcotuli, I.; Blanco, A.; Maccaferri, M.; Condorelli, G.E.; Tuberosa, R.; Parada, R.; de Camargo, A.C.; Schwember, A.R.; Gadaleta, A. Carotenoid pigment content in durum wheat (Triticum turgidum L. var durum): An overview of quantitative trait loci and candidate genes. Front. Plant Sci. 2019, 10, 1347. [Google Scholar] [CrossRef] [Green Version]
  196. Patil, R.; Oak, M.; Deshpande, A.; Tamhankar, S. Development of a robust marker for Psy-1 homoeologs and its application in improvement of yellow pigment content in durum wheat. Mol. Breed. 2018, 38, 136. [Google Scholar] [CrossRef]
  197. Singh, R.P.; Singh, P.K.; Rutkoski, J.; Hodson, D.P.; He, X.; Jørgensen, L.N.; Hovmøller, M.S.; Huerta-Espino, J. Disease impact on wheat yield potential and prospects of genetic control. Annu. Rev. Phytopathol. 2016, 54, 303–322. [Google Scholar] [CrossRef] [Green Version]
  198. David, B.G. Hessian Fly (Diptera: Cecidomyiidae) Injury and Loss of Winter Wheat Grain Yield and Quality. J. Econ. Entomol. 1999, 92, 1190–1197. [Google Scholar]
  199. Raman, H.; Ryan, P.R.; Raman, R.; Stodart, B.J.; Zhang, K.; Martin, P.; Wood, R.; Sasaki, T.; Yamamoto, Y.; Mackay, M. Analysis of TaALMT1 traces the transmission of aluminum resistance in cultivated common wheat (Triticum aestivum L.). Theor. Appl. Genet. 2008, 116, 343–354. [Google Scholar] [CrossRef] [PubMed]
  200. Luo, M.-C.; Dvořák, J. Molecular mapping of an aluminum tolerance locus on chromosome 4D of Chinese Spring wheat. Euphytica 1996, 91, 31–35. [Google Scholar] [CrossRef]
  201. Cabanillas, B. Gluten-related disorders: Celiac disease, wheat allergy, and nonceliac gluten sensitivity. Crit. Rev. Food Sci. Nutr. 2020, 60, 2606–2621. [Google Scholar] [CrossRef] [PubMed]
  202. Sharma, N.; Bhatia, S.; Chunduri, V.; Kaur, S.; Sharma, S.; Kapoor, P.; Kumari, A.; Garg, M. Pathogenesis of celiac disease and other gluten related disorders in wheat and strategies for mitigating them. Front. Nutr. 2020, 7, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Scherf, K.A. Immunoreactive cereal proteins in wheat allergy, non-celiac gluten/wheat sensitivity (NCGS) and celiac disease. Curr. Opin. Food Sci. 2019, 25, 35–41. [Google Scholar] [CrossRef]
  204. Sabença, C.; Ribeiro, M.; Sousa, T.d.; Poeta, P.; Bagulho, A.S.; Igrejas, G. Wheat/Gluten-Related Disorders and Gluten-Free Diet Misconceptions: A Review. Foods 2021, 10, 1765. [Google Scholar] [CrossRef] [PubMed]
  205. García-Molina, M.D.; Giménez, M.J.; Sánchez-León, S.; Barro, F. Gluten free wheat: Are we there? Nutrients 2019, 11, 487. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Barro, F.; Iehisa, J.C.; Giménez, M.J.; García-Molina, M.D.; Ozuna, C.V.; Comino, I.; Sousa, C.; Gil-Humanes, J. Targeting of prolamins by RNAi in bread wheat: Effectiveness of seven silencing-fragment combinations for obtaining lines devoid of coeliac disease epitopes from highly immunogenic gliadins. Plant Biotechnol. J. 2016, 14, 986–996. [Google Scholar] [CrossRef] [Green Version]
  207. Gil-Humanes, J.; Pistón, F.; Tollefsen, S.; Sollid, L.M.; Barro, F. Effective shutdown in the expression of celiac disease-related wheat gliadin T-cell epitopes by RNA interference. Proc. Natl. Acad. Sci. USA 2010, 107, 17023–17028. [Google Scholar] [CrossRef] [Green Version]
  208. Henggeler, J.C.; Veríssimo, M.; Ramos, F. Non-coeliac gluten sensitivity: A review of the literature. Trends Food Sci. Technol. 2017, 66, 84–92. [Google Scholar] [CrossRef]
  209. Pastorello, E.A.; Farioli, L.; Conti, A.; Pravettoni, V.; Bonomi, S.; Iametti, S.; Fortunato, D.; Scibilia, J.; Bindslev-Jensen, C.; Ballmer-Weber, B. Wheat IgE-mediated food allergy in European patients: α-amylase inhibitors, lipid transfer proteins and low-molecular-weight glutenins. Int. Arch. Allergy Immunol. 2007, 144, 10–22. [Google Scholar] [CrossRef] [PubMed]
  210. Plamenov, D.; Belchev, I.; Spetsov, P. Anther culture response of Triticum durum × T. monococcum ssp. aegilopoides amphiploid. Cereal Res Commun. 2009, 37, 255–259. [Google Scholar] [CrossRef]
  211. Li, H.; Liu, X.; Zhang, M.; Feng, Z.; Zhang, L. Development and identification of new synthetic T. turgidumT. monococcum amphiploids. Plant Genet. Resour. 2018, 16, 555–563. [Google Scholar] [CrossRef]
  212. Gul Kazi, A.; Rasheed, A.; Bux, H.; Napar, A.A.; Ali, A.; Mujeeb-Kazi, A. Cytological, Phenological and Molecular Characterization of B (S)-Genome Synthetic Hexaploids (2n = 6x = 42; AABBSS). Cereal Res. Commun. 2015, 43, 179–188. [Google Scholar] [CrossRef] [Green Version]
  213. Rafique, K.; Rasheed, A.; Kazi, A.G.; Bux, H.; Naz, F.; Mahmood, T.; Mujeeb-Kazi, A. Powdery mildew resistance in some new wheat amphiploids (2n = 6x = 42) derived from A- and S-genome diploid progenitors. Plant Genet. Resour. 2012, 10, 165–170. [Google Scholar] [CrossRef]
  214. Ishii, H. A quantitative study of morphological relationships among natural tetraploid wheats and their putative ancestors. Sci Bull. Coll. Agric. Univ. Ryukyus 1972, 19, 75–85. [Google Scholar]
  215. Wang, K.; Shi, L.; Liang, X.; Zhao, P.; Wang, W.; Liu, J.; Chang, Y.; Hiei, Y.; Yanagihara, C.; Du, L.; et al. The gene TaWOX5 overcomes genotype dependency in wheat genetic transformation. Nat. Plants 2022. [Google Scholar] [CrossRef]
  216. Luo, J.; Li, S.; Xu, J.; Yan, L.; Ma, Y.; Xia, L. Pyramiding favorable alleles in an elite wheat variety in one generation by CRISPR-Cas9-mediated multiplex gene editing. Mol. Plant 2021, 14, 847–850. [Google Scholar] [CrossRef]
Figure 1. Evolution of Triticum turgidum and Triticum timophevvi group wheat. BP, before present. The dashed gray arrow denotes the possible but not confirmed pathway for the origin of bread wheat (T. aestivum L.).
Figure 1. Evolution of Triticum turgidum and Triticum timophevvi group wheat. BP, before present. The dashed gray arrow denotes the possible but not confirmed pathway for the origin of bread wheat (T. aestivum L.).
Ijms 23 02206 g001
Table 1. Botanical category of tetraploid wheat.
Table 1. Botanical category of tetraploid wheat.
PloidyCategoriesBotanical NameSynonymsPhenotype
Tetraploid
2n = 28
T. timopheevi Zhuk. AuAuBspBsp/ AuAuGGvar. araraticumT. araraticum Jakubz.wild, brittle rachis, hulled
concv. timopheeviT. timopheevi Zhuk.domesticated, hulled
T. turgidum L.
AuAuBB
var. dicoccoidesT. dicoccoides Koern.wild, hulled, brittle rachis
concv. dicoccon/dicoccumT. dicoccon Schrank.domesticated, hulled,
semi-brittle rachis
concv. durumT. durum Desf.cultivated, free-threshing,
tough rachis
concv. turgidumT. turgidum L.cultivated, free-threshing,
tough rachis
concv. polonicumT. polonicum L.cultivated, free-threshing,
tough rachis
concv. carthlicumT. carthlicum Nevski.cultivated, free-threshing,
tough rachis
concv., cultivar-group; var., variety.
Table 2. Summary of the improvement of durum wheat through diversity germplasms and approaches.
Table 2. Summary of the improvement of durum wheat through diversity germplasms and approaches.
TraitGenes/QTLsChromosomeApproachesOriginVariationsReferences
Yield componentplant heightRht1,
Rht14,
Rht15
4B,
4A,
6A
pentaploid hybridization,
homologous hybridization
hexaploid wheat Norin10,
durum wheat Castelporziano or Durox
reduced height; increased harvest index or pleiotropic effect[145,146,147]
multiple traits//pentaploid hybridizationhexaploid wheat CSCR6pleiotropic effect[148]
/4Dpentaploid hybridizationAe. tauschii accession AT23big spike, significantly increased number of spikelets and florets per spike;
enhanced YR resistance
[149]
grain sizeGW2-A1,
GW2-B1
6AS,
6BS
RNA interferencedurum wheat Svevoincreased kernel size[150]
Grain qualitysoft kernelPina,
Pinb
5Dph1c-mediated
homoeologous recombination
Langdon 5D(5B)
substitution line,
durum wheat ph1c line Cappelli M
soft grain, vitreous kernels, high GPC, and good gluten quality[151]
ph1b-mediated
homoeologous recombination,
homologous recombination
hexaploid wheat CS,
Langdon-CS 5D(5B)
substitution line
soft grain[152]
Pina,
HMW glutenin subunit 1A × 1
1D,
1A
gene editingdurum wheatsoft kernel and better breadmaking quality[153]
flour propertiesGlu-D11Dph1-mediated
homoeologous recombination
multiple germplasmsimproved bread-making quality[28,154,155,156,157,158,159]
grain protein contentGpc-B16Bhomologous recombinationLangdon-T. dicoccoides (6B) substitution lineincreased protein level[160]
yellow pigment concentrationLCYe,
HYD2
Targeting Induced Local Lesions in Genomes (TILLING)durum wheatsignificantly increased β-carotene accumulation[161]
pre-harvest sproutingQphs.sicau-3B.13Bpentaploid hybridizationhexaploid wheat
T. spelta CSCR6
high resistance to PHS[162]
Biotic resistanceleaf rust,
powdery mildew,
tan spot
Pm13,
Lr19,
YPC
3B,
7A
ph1 mediated
homoeologous recombination
Ae. Longissima,
Agropyron
improved resistance to PM, LR, and YPC[156,163,164]
multiple/inter/intra-specific hybridizationT. araraticum,
T. dicoccoides,
Ae. speltoides
enhanced resistance to LR and tan spot;
increased grain yield and GPC
[165]
Lr34/Yr18/Sr57/Pm38/Ltn1/gene editingbread wheatrobust seedling resistance to LR, YR, and PM[166]
stem rustSr227Ainterspecific hybridizationT. monococcum L. cv. RL 5244differential resistance depends on ploidy level[167]
stripe rust//interspecific hybridizationA-genome diploidsenhanced resistance to YR, LR, and PM[168,169,170]
Yr284Dpentaploid hybridizationAe. tauschii accession AT23enhanced resistance to YR[149]
fusarium head blightFhb13BS,
4AL,
4BS,
5AL,
6AS
homologous recombinationhexaploid wheat
Sumai-3
improved resistance to FHB[171]
Qfhb.ndwp-5A, Qfhb.ndwp-7A5A,7Ahomologous recombinationhexaploid wheat
PI 277012
[172]
/2AS,
2BS,
3AL,
4BL
pentaploid hybridizationhexaploid wheat
Sumai-3
[173]
//pentaploid hybridizationhexaploid wheat
Sumai-3
[174]
//mutation by treating with DNA methylation inhibitor (5-methyl-azacytidine)/[175]
Hessian fly///durum wheat
resistant lines
enhanced resistance to Hessian fly,
superior agro-phenological traits
[176]
Abiotic resistancethermotoleranceheat shock proteins/gene modify (overexpression)/improved thermotolerance[177,178]
water//intraspecific hybridizationwild emmerenhanced adaptation to water stress[179]
drought,
salt
TdPIP2;1/gene modify (overexpression)durum wheatenhanced drought and salt tolerance[180]
salt//induced by ascorbic acid/enhanced salt tolerance[181]
Kna1 locus, TaALMT1 gene4Dpentaploid hybridization, ph1-mediated
homoeologous hybridization, homologous hybridization
bread wheat or Langdon-CS 4D (4B) substitution lineenhanced Al3+ tolerance[182,183,184]
//intraspecific hybridizationT. emmer,
T. durum,
T. carthlicum,
T. turgidum,
T. turanicum,
T. polonicum
enhanced salt tolerance[185,186]
wheat-related disorderscoeliac diseaseα-gliadin genes/CRISPR/Cas9 (knock out)durum wheathighest 69% reduction of gliadin of reduced immunoreactivity[187]
onset of bakers’ asthmaWTAI-CM3, WTAI-CM16/CRISPR/Cas9 (knock out)durum wheatreduced allergen proteins[188]
/, unspecified; YPC, yellow pigment concentration; FHB, fusarium head blight; PM, powdery mildew; LR, leaf rust; SR, stem rust; YR, stripe rust; PHS, pre-harvest sprouting; LCYe, lycopene ε-cyclase; HYD2, β-carotene hydroxylase 2; HMW, high molecular weight; GW2, grain weight 2; WTAI, wheat α-amylase/trypsin inhibitors; CS, Chinese Spring.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yang, F.; Zhang, J.; Liu, Q.; Liu, H.; Zhou, Y.; Yang, W.; Ma, W. Improvement and Re-Evolution of Tetraploid Wheat for Global Environmental Challenge and Diversity Consumption Demand. Int. J. Mol. Sci. 2022, 23, 2206. https://doi.org/10.3390/ijms23042206

AMA Style

Yang F, Zhang J, Liu Q, Liu H, Zhou Y, Yang W, Ma W. Improvement and Re-Evolution of Tetraploid Wheat for Global Environmental Challenge and Diversity Consumption Demand. International Journal of Molecular Sciences. 2022; 23(4):2206. https://doi.org/10.3390/ijms23042206

Chicago/Turabian Style

Yang, Fan, Jingjuan Zhang, Qier Liu, Hang Liu, Yonghong Zhou, Wuyun Yang, and Wujun Ma. 2022. "Improvement and Re-Evolution of Tetraploid Wheat for Global Environmental Challenge and Diversity Consumption Demand" International Journal of Molecular Sciences 23, no. 4: 2206. https://doi.org/10.3390/ijms23042206

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop